Principles of Chemical Science MIT Lecture Videos
Videos 21-30 of 32
 

Lecture on acid base titration
The video features a lecture on acid base titration delivered by Professor Catherine Drennan. In her lecture Professor Drennan explained the principle of acid base titration and procedure of calculations. Edited by Ashraf
Tags // Lecture on acid base titration
Added: April 16, 2009, 11:25 pm
Runtime: 3004.47 | Views: 21640 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Video lecture: oxidation reduction
The video presents a lecture on oxidation-reduction. Oxidation-reduction is important for many biological processes. Edited by Ashraf
Tags // Oxidation Reduction
Added: April 16, 2009, 11:26 pm
Runtime: 2791.73 | Views: 12433 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - Oxidation Reduction (cont.) - Lecture 26
Principles of Chemical Science/n * Email this page/nVideo Lectures - Lecture 26/nTopics covered: /nOxidation/Reduction (cont.)/nInstructor: /nProf. Catherine Drennan/nTranscript - Lecture 26/nGood afternoon./nGood afternoon everybody. I am going to start this morning, or this afternoon with a question we had at the bottom of last lecture's handout./nBut we are not going to actually answer this question until the end of today's lecture./nLet's look at what the question is, and you can be thinking about it during today's lecture. All right. The question is vitamin B12, which we have already mentioned once in this course, it has a large negative redox potential./nThe standard reduction potential is large and negative./nHow does it get reduced in the body? Let's think about the problem here. As we learned last time, compounds with large negative reduction potentials are hard to reduce./nThey are easy to oxidize, which makes them good reducing agents, but they are hard to reduce. And vitamin B12 has a large negative value so it is hard to reduce./nBut in the body it needs to be reduced to be active for one particular enzyme./nIt needs to be reduced but it is hard to reduce, so what does the body do to compensate? How does this work? I am going to answer this at the end of class, but let me first tell you in the beginning of class why this is an important question./nThe proper functioning of a particular enzyme requires reduced B12 and also another B vitamin called folic acid./nAnd this particular enzyme is believed to be important in preventing heart disease and also important in preventing birth defects and also important for maintaining mental health./nThis September, Laura Philips, who is a recitation instructor, and I went to Oxford to attend a meeting on vitamin B12. And it was a very medical-based meeting./nAnd a large number of the talks were by physicians talking about treating people with B12 deficiencies that had mental illnesses./nAnd they told one story of a man who was sort of a middle-aged man who couldn't really function at work. He couldn't remember what he was supposed to be doing, and so he was having trouble. He was having trouble driving./nHe couldn't remember all the rules of the road. And of course, it seems like in England, this guy was in England, that was even more complicated./nAnd he would leave his house and couldn't remember where he lived./nHe had all these symptoms that were sort of Alzheimer's disease but he was not an older man. And it did not seem like Alzheimer's was right so eventually he ended up with this physician who diagnosed that he had a B12 deficiency./nAnd it was great for this particular individual because when they started giving him vitamin B12 in high doses he actually recovered. He got his driver's license back./nHe got his job back. And he was able to get back and forth from home without any challenges./nI think a lot of the time one cannot point to B12 as the source of any mental forgetfulness. And people do ask me. People advertise this. Like in health food stores you can buy vitamin B12 and it talks about improving one's memory and things like that./nI think that if you had a vitamin B12 deficiency you should be taking B12./nBut I think if you do not have a B12 deficiency taking excess is not going to help you do better on exams. Although, I guess they do have squirt B12, just a little extra dose squirted in the nose before the exam just to be on the safe side in case that helps./nI won't vouge for anything like that with memory and B12, but nonetheless these enzymes are important. And it is important to figure out how the B12 gets reduced so the enzyme can function. What is the enzyme?/nThe enzyme's name is methionine synthase, and it uses these two B vitamins, vitamin 12 and folic acid./nIt converts an amino acid called homocysteine to another one call methionine. And it also converts one form of the folic acid to another form of the folic acid. And, as I said, inhibition of this enzyme is believed to be involved with heart disease, it has been linked to birth defects and also anemia./nAnd perhaps a real deficiency has been linked to mental issues./nThis is a pretty important enzyme, but there are only two enzymes in humans that use vitamin B12. Birth defects in one of those enzymes can lead to this problem I talked about with buffering the blood./nAnd that can lead to death in infants if it is not corrected, if it is too severe of a problem because the blood cannot handle the change in pH. And this one is also quite important medically. At this point, I think I may have scared you that you want to know if you are getting enough B12 and folic acid in your diet./nSo I thought I would share with you where you get B12 and folic acid in your diet./nAt the end of class we will talk about how B12 is reduced. Where do you get these in your diet? Well, for vitamin B12, does anybody know, in your diet? I have heard spinach. No./nTotal cereal that is supplemented with vitamins, that is a good way to get it./nBut naturally occurring it comes in meat, red meat and other forms of meat. This is the one vitamin that vegetarians say they need to have supplemented. And it is not a huge problem because, as I said, cereal and other things./nBut in the natural diet it comes from meat. And what about folic acid?/nVegetables, yeah. Folic from foliage, so green leafy things work. Also I was at another meeting with medical doctors talking about vitamin B12 and folic acid./nThese were Norwegian physicians. They informed me of another source of folic acid that I was unaware of, which is apparently Norwegian beer./nI cannot vouge for this but that is what they told me./nAlso perhaps a better source would be leafy green vegetables. And orange juice is actually quite good. Some of you may have seen this commercial, raise your hand if you have seen this, with orange juice and it says orange juice, it is good for your heart? Have you seen that ad? That actually is probably true./nAnd that is true because you get a lot of folic acid in orange juice./nAnyway, here are some dietary suggestions for you. All right. But you actually don't have to worry too much about folic acid anymore because now it is fortified in bread. In addition to breakfast cereals, bread now has extra folic acid./nAnd this started as a campaign because if you are a woman who is pregnant you need a lot of folic acid or the baby can be born with birth defects./nNow, the problem is that often you need the folic acid before you know that you are pregnant./nYou need the folic acid from day one. And so a lot of scientists have been campaigning in Washington for a long time saying we need to put the folic acid in bread because the woman need the folic acid before they know that they are pregnant./nThis argument did not go very well, and it was years and years and years of this argument./nAnd then all of a sudden some studies came out that really linked folic acid and heart disease through the activity of this B12 dependent enzyme, and all of a sudden the climate in Washington changed and they say, oh, this may prevent heart disease./nAll right. Let's fortify bread. That is a good example of sort of science and politics where political decisions that are made about science are really based on politics and not on the science, but at this point things are looking better./nAlthough, some recent studies have shown there are still issues for some women in getting enough folic acid./nAll right. Then we need to know how it is reduced. We are going to come back to this at the end of today and look at how all the things we have been learning about oxidation-reduction apply to not only the batteries that you buy but to the reactions such as the reaction of methionine synthase, which is actually happening in your body right now./nAll right./nToday we are going to talk about, this final lecture of oxidation and reduction, adding and subtracting half cell reactions. This is listed as a topic, but I decided to remove this so it is not actually in your handout, although it is still listed as being in your handout./nYou can cross that out just for interest of time. And we are going to talk about the Nernst equation, which is one of my favorite equations, so I am looking forward to that. All right./nA couple other announcements before we get into this material./nI want to remind people there is a forum tonight. We had a great forum last week, and so I encourage you to sign up and come to these forums. Also the problem set is due this Friday, as usual. And then the next problem set is a really short problem set that will be due Wednesday./nThat problem set just covers material on Friday's class and Monday's class, so it is actually really very short. OK./nDown to some business. All right. Adding and subtracting half cell reactions./nSuppose you needed to know something about this particular equation and the standard reduction potential for the reaction of copper two with one electron going to copper plus one, but it wasn't given and you are looking in the table and it is not given./nHowever, other things related are given./nSuppose you do find in the table copper plus two with two electrons going to copper zero, copper solid, and you find copper solid going to copper plus one plus one electron./nThat information about both of those reactions is provided, so you see the relationship between these two equations. You can use the information about these to calculate a standard reduction potential for this one./nIf you add these two reactions together you get the reaction of interest./nYou get copper two and one electron going to copper one, one electron cancels and the copper solid cancels and you get the reaction of interest./nIf you can add together two reactions to get the reaction of interest, what do you do with the standard reduction potentials to get the answer for the reduction potential for this reaction? This all comes back to delta G./nMost things are related in some ways to delta G, free energy. The free energy in the standard state, delta G knot, for that new reaction that you are interested in is equal to the delta G knot for the reaction that was a reduction in the added equations minus the delta G information for the oxidation reaction./nYou notice we are doing a derivation here? All right. Then we know something about the relationship between delta G and standard reduction potentials./nWe talked about this last time that delta G knot equals minus N, which is the moles of electrons times Faraday's constant times standard reduction potential./nIf we substitute that into all of these we get this equation, and we can cancel out the Faraday's constant and solve then for the standard reduction potential for the new reaction, which is what we want./nIf we get rid of Faraday's constant it cancels out./nBring the number of moles of electrons for the new reaction down here. We get this equation which tells you about the number of moles of the reaction in the standard reduction potential for the reduction reaction, number of moles of electron, the standard reduction potential for the reaction that is in oxidation./nThat is added to the reduction to give you this new reaction./nAnd that is divided by the number of moles electrons involved in that new reaction that you are interested in. Let's use that then. If we go and we look, these were, are two reactions that we are adding together./nOne is an oxidation. One is a reduction. And we can look up what the standard reduction potentials are that are related to those two equations. The standard reduction potential, copper two to copper, copper plus one to copper./nAgain, we are looking up reduction potentials./nAnd we can find out what those are and then we can plug them in. Here is our equation again. For the reduction, we have two electrons involved in the reduction and a standard reduction potential of plus 0.340 volts./nThen we put in the information about the reaction that was written as an oxidation./nAnd so that is minus. There is one electron involved in that. And a reduction potential. Again, it is entered in as the reduction potential./nAccording to the equation this is a reduction potential of 0.522 volts. And then, in the final reaction, there is one electron involved in that. And so we put the one here and we can get out an answer of plus 0.158 volts./nAnd so now we found out a new standard reduction potential for this half reaction knowing information about other half reactions./nAnd so this would be then the copper two cooper one reduction potential. What about the cells we were talking about before, the electrochemical cells with the balanced equations?/nWell, in that case, for an electrochemical cell, the number of moles of electrons released at the anode is going to be equal to the moles taken up at the cathode, which is equal to the overall moles that get balanced out./nWhen you balance the equation, you always want to cancel out your electrons and have none left in your balance equation. Otherwise, it is not balanced. In that case, you don't use this./nIt simplifies to the equation that we talked about before where the reduction potential for the cell now, it is the cell potential, not another half potential, is equal to the reduction potential for the reaction at the cathode minus the reduction potential for the reaction at the anode./nYou use this for balanced cells and you use this for half cell reactions. And so, again, for a half cell reaction you must use this. Otherwise, it is not going to work out. You can recognize a half cell reaction because it still has electrons in it./nThat is a half cell reaction./nOK. That is just sort of a little side thing that will come in handy in finishing up the problem set and on the test. And now we can talk about one of my more favorite topics, which is the Nernst equation./nA lot of you have probably experienced the ramifications of the Nernst equation without realizing that is what you have done. I assume that pretty much everybody has had an occasion where you to use your favorite gadget and discover that the battery is no longer functioning./nA few people say yeah, maybe that has happened to me./nAnd so then you start doing things like this where you either get some kind of recharger or you go out and stockpile batteries, which then if you keep them long enough it seems like they don't work either./nAnyway, that is all the Nernst equation in action. An exhausted battery is a sign that equilibrium has been reached./nAt equilibrium the cell generates zero potential across its electrodes, so it is not working anymore./nAnd so to understand this problem, we need to think about how the composition of that cell is changing over time. This gets back to all the other things that we have talked about with chemical equilibrium./nWhat do we know about this kind of stuff?/nWhat do we know about equilibrium and components of the reaction? We know that delta G changes as the composition of the reaction changes, and we know that it changes until we reach at equilibrium delta G equals zero./nWe know that delta G at some given time, at some different reaction mixture is equal to delta G knot, the delta G at standard conditions./nPlus RT, the gas constant times temperature, times the natural log of Q./nAnd Q is what? Reaction quotient. The delta G at any particular time depends on the composition, depends on the reaction that is going on, the components of that reaction. We have seen this before./nLast time we talked about the relationship between delta G and cell potential, delta E, so we know something about that relationship./nWe know that delta G knot is equal to the number of moles of electrons times Faraday's constant times delta E knot, the change in cell potential for the reaction./nAnd also this could be applied if this was delta G to delta E./nIt holds if you are under standard conditions, delta G knot, delta E knot, or under nonstandard conditions. If we combine those things, we get the Nernst equation. If we take what we already knew about delta G from before and now plug in, instead of delta G knot, minus N Faraday's constant times delta E knot./nAnd also we are going to substitute in for delta G minus N Faraday's constant times delta E over here, and we substitute over here and we keep this part the same./nAnd then if we divide by N and Faraday's constant, we get the Nernst equation here./nThe delta E at some given point for some different reaction quotient at some different time in the cell is going to be equal to the standard cell potential under standard conditions minus RT gas constant temperature over the number of moles of electrons times Faraday's constant times the natural log of Q, the reaction quotient./nWhat the cell potential is, at some given time, depends on the reaction. Where you are in the reaction./nWhat the composition of products to reactants is at that given time. This way we can calculate how close the battery is to being dead, sort of where we are in the trajectory of things, what kind of potential we are getting at that particular given time./nLet's do an example of doing the Nernst equation./nHere is a particular reaction. We have copper plus two plus zinc solid going to zinc plus two plus copper. And we are told that the zinc ion's concentration is 0.10 molar and copper plus two is 0.0010 molar./nAnd so these are not standard conditions so we know we want to use the Nernst equation./nThe first step in doing this is we will write down the Nernst equation. The cell potential at some given time is equal to the standard cell potential minus RT over N times Faraday's constant times the natural log of Q./nWe will have step one being to find the standard cell potential, so find delta E knot./nAll right. To do this we need to look at both the reactions. We need to know the standard reduction potential for both reactions./nOne reaction involves copper plus two going to copper solid. And this reaction involves two electrons. The other reaction that is going in the forward direction is that zinc solid is going to zinc plus two plus two electrons./nWe have a reduction and an oxidation./nThis reaction would be occurring where, at the cathode or the anode? At the cathode. And so that must mean that the other one is happening at the anode./nNow we need to know the standard reduction potential for the two couples./nWe need to know the reduction potential of copper two plus to copper solid, and we need to know the potential for zinc two plus to zinc solid. And we can look those up in a table. They are up here, if you can see them./nI am actually going to write them down to a couple more significant figures./nOne is 0.340 and the other is minus 0.7628, and those are in volts. OK. We have that information. And now this is a balanced reaction. There were no extra electrons floating around./nThis is a cell./nWe are trying to find a cell potential here. We can calculate that for the particular cell in question from the standard reduction potential for the reaction at the cathode minus the standard reduction potential for the reaction at the anode./nAnd so that is going to be equal to 0.34 minus a negative 0.7628./nWhich is going to give you 1.103 volts. This is the reaction we use. And, again, we are not changing the sign because we are looking up and entering standard reduction potentials./nBut you need to know which is the reaction at the cathode and which is the reaction at the anode to put them in the right place./nAll right. We have our first thing there. Now what do we need to know? Well, we also need to know Q./nStep 2, let's find Q. What is Q going to be equal to? What is on top? Concentration of?/nYeah, zinc two plus and nothing else because it's products over reactants. But we don't include solids in our equation, so it's just zinc two plus./nAnd that is going to be then over copper two plus. Again, we are not including solids. We are not including the zinc solids. Products over reactants, and we can plug this in./nWe have 0.1 over 0.0010, and that gives us 1.0 times 10 to the 2 for Q./nStep 3, we also need to know N. How many moles of electrons are involved in this reaction? Two. N equals two. We have from a copper two plus to copper solid zinc two plus to zinc solid./nThere are two electrons evolved./nThey are not in the equation because they canceled out. This one is obvious. Sometimes they are harder and you actually really have to think about how that equation was balanced to come up with the right answer for the number of moles./nAll right./nNow we can plug all that information in to the Nernst equation./nWe are looking for what the cell potential is at a particular time, so that will be equal to the standard cell potential that we calculated minus RT./nAnd R is the gas constant, so write the gas constant out, times the temperature. And the temperature for all the problems that you are going to do in this unit is always going to be at room temperature./nTimes the natural log of Q./nAnd that is going to be over two times Faraday's constant. And I am writing all this out because we are going to talk about units in a minute. Times Faraday's constant. Now we can cancel out. We will cancel moles./nAnd we are going to cancel Kelvin./nAnd we are going to left with joules per Coulomb. Is that a good thing? What does joules per Coulomb equal? A volt, right. That will do very well because we are going to have a volt here and we are going to subtract something in volts./nThat will leave us with joules and a volt equals joules per Coulomb./nThat gives us 1.103 volts minus. This calculates out to 0.0592 volts. And, if we worked this out, this would give us a number with two significant figures after the decimal point./nAnd so that leaves us with these significant figures here. We are actually going to lose one in the subtraction anyway. And we come up with an answer of 1.044 volts./nI will mention, in terms of significant figures, this is like the nastiest problem./nYeah? Do you have a question?/nSometimes when you look at the equation there is a lot of different things going on, and you actually have to do sort of a step backwards to think what were the half reactions and what was canceled./nA lot of times it is very obvious, but some of them are a little bit longer./nThere may be one in the problem set, I don't remember, but you will know what I mean when you see it. Sometimes it will be just very obvious like that, that it will be two. Other times it is a little bit harder to see, and you sort of want to think backwards to what those half reactions were and how you would have balanced them./nAll right. In terms of significant figures, these problems are probably some of the worst in terms of significant figures./nAcid base is pretty bad, too. You notice we started with a subtraction, and in the subtraction you often lose significant figures./nThen we had a division. Then we had a log. And then we had multiplication and division and then another subtraction. The only thing I can say for these problems is that it helps to really write out every step and to put a line under where your significant figure is after each calculation to make sure that you aren't getting the wrong significant figures at the end./nYou cannot, in this case, go back and say how many significant figures did I have in the information that was given to me? That is how many I have at the end./nThat is not going to work for you. These are a little bit tricky. A lot of the time you will have a break because there won't be many significant figures in this answer. And so when you do the subtraction you kind of compensate./nSometimes it doesn't matter how many you really have here because you are going to lose them in that subtraction step anyway, but just to warn you there is a lot of significant figure stuff./nThis actually would be a great significant problem if that all you were testing because you have to do every kind of significant figure situation in this one problem, but I am going to make your life slightly easier./nWhat I am going to do is do some of the steps for you./nIf we talk about constants for a minute, we are always going to be at room temperature in this unit because that is just where we are. It is going to make it harder to anything else. Unless specified, assume room temperature./nIf you do that you will find that this is always going to be the same number. You have the gas constant, the temperature and Faraday's constant, and so there is one number that that is equal to./nAnd so you can use this number./nOr, if you prefer to use log base 10 rather than natural log, you can take this number and multiply it by the conversation factor and you have sort of another constant which you can use in the problem./nAnd, on the test, I will actually give the equations in this format to you on your sheet./nAnd so you can pick one of these. If there is room I will give both. Otherwise, I will just give one. But it is the same type of problem./nYou just have to pay attention to which log to use. And I will make sure that the number of significant figures here is not limiting in the problem, so you don't have to worry about that. I highly recommend that if I provide these two you, save those extra steps of multiplying things out./nIt will save you some time and maybe a little bit of significant figure headache./nThose will be available just to simplify these problems. I am sure that you can all multiply RT over Faraday's constants, so I am just going to provide that for you and not require you to do it on the test./nWhat happens at equilibrium then? At equilibrium what does Q equal? Q equals K at equilibrium, right./nWhat about delta G? What does delta G equal at equilibrium? Zero. Do you see where we are going with this? If you have the handout you can look ahead./nWe know from before that delta G equals delta G knot plus RT natural log of Q./nWe also know that delta G knot equals minus RT natural log of K, because at equilibrium delta G is zero so we drop that out and add equilibrium./nQ equals K. We have seen this expression before so we know that. We also know the relationship between delta G and cell potential. We know that delta G knot equals minus N, the number of moles of electrons./nFaraday's constant times the standard cell potential./nNow we can rearrange these and relate the equilibrium constant to cell potential. Combining we get this, so we just set those two equal to each other and we can solve for K, or at least natural log of K in terms of cell potential./nAnd so you can calculate the equilibrium constant from cell potentials./nAnd, of course, this is important because if you want your cell to be producing some kind of current creating energy or if you want to apply something to get a non-spontaneous reaction to go, you care about what the composition of the reaction is, you care about the equilibrium constant and you care about these relationships./nIf you know about K, you can also think about if the reaction is really favorable in one direction or in the other direction at equilibrium do you have a lot more products or a lot more reactants, some of the same questions we have talked about before./nSee how all these units are kind of fitting together and how I told you that everything that you learned you have to remember because it is going to come back? It's back again. We are back to thinking about equilibrium constants./nThis half semester of the course is very, very connected./nYou cannot get away from any of the material, which is kind of good because on the exam it is kind of all related, so doing the next unit is reviewing the unit that you already did. All right. Let's look at an example of an oxidation-reduction reaction and talk about what the equilibrium constant is again at room temperature./nThis is going to involve two half reactions./nWe have a reaction with zinc in the forward direction, I mean with lead plus two going to lead solid. Lead plus two electrons going to lead solid. We also have zinc solid going to zinc plus two and then two electrons./nThose are the two half reactions involved in this cell reaction. Now, we are interested in K./nWe are interested in the magnitude of K. We want to think about if K is big. Is K greater than one for this reaction? We can do the math and calculate, knowing something about the cell potentials, or we can just think about whether this reaction would be favorable before doing the math./nIf you think about what is happening in the forward reaction, what is happening? Well, what is happening is that zinc is reducing lead plus two./nLead plus two is being reduced to lead solid and zinc is being oxidized./nIt is acting as a reducing agent. And in the backward reaction what is happening? Well, in the backward reaction we have lead solid reducing zinc plus two to zinc solid. If we think about are there going to be more products or reactants, we are really thinking about which of these two things is a better reducing agent./nIf zinc is a better reducing agent than lead it will favor the forward direction and we will have a K greater than one./nIs it a better reducing agent? We talked last time about how to figure this out, and we talked last time about the fact that things with positive numbers up here, the oxidized species of this is strongly oxidizing./nAnd at the bottom with these negative numbers, the reduced form is strongly reducing./nWe can look at where things are. There are numbers, there are standard reduction potentials, and figure out which one is a better reducing agent. If we look those up, we know the zinc couple. We just talked about that one, actually, minus 0.7628./nAnd we can look up the standard reduction potential of lead, and it is also negative but it is a smaller negative number minus 0.1263./nThis is a big negative number. Is it going to be a good reducing agent? What do we know about how difficult it is to add electrons to this? Is it hard to add electrons to this? Yeah./nIt is hard to reduce. That makes it easier to oxidize. Does that make it a good reducing agent? Yeah, it does./nBecause a reducing agent gets oxidized itself. It reduces other things. This would be a good reducing agent./nWell, what about this? Is it as good? No, it is not as good. It is probably not bad, but it is not as good a reducing agent. It would be a little easier to reduce than zinc, a little harder to oxidize, so it is not as good a reducing agent because it is a more positive number than this one./nThis is the bigger negative number so this is the better reducing agent./nIf that is the better reducing agent then the equilibrium should lie to the right. There should be more products than reactants at equilibrium because in the forward direction zinc is acting as the reducing agent./nAnd it is a better reducing agent so it should drive the equilibrium such that there are more products than reactants. And so K should be greater than one./nAnd we can actually do the math and prove that this is true./nThis is very much like when we talked in acids and bases about equilibrium and K values and how if something was a stronger acid one way then a stronger acid in the backward reaction you could predict what K would be or you could do the math./nIt is the same principle here. If we go through the math, we can plug in the standard reduction potential for the lead couple and the standard reduction potential for the zinc couple and get an answer for the cell potential./nWhat is N for this? This is also a fairly simple one? Two./nWe know N is two. And then we can plug into our equation. That is all we really needed to know to do this one. We needed to know the standard potential for the reaction, which we calculated, Faraday's constant, we need to N and RT./nAnd so we can put those in./nAnd I will also provide this with the math worked out for Faraday's constant and R and T. And so we can plug that in and put in our two and put in the number of volts, the standard cell potential that we calculated just before and go back and calculate for K./nAnd so K is 21.5. That is definitely bigger than one./nThat is a very big number. Sorry, log of K and K turns out to be a very, very big number. And so the forward reaction is definitely favored, which is what we assumed thinking about which was a better reducing agent./nIt said that the forward reaction should be favored, that K should be greater than one. And when you do the math you find that K is, in fact, greater than one. It is favored in that direction. More products than reactants at equilibrium./nIn the last couple of minutes then I want to go back to biological systems and show you that the same equations apply./nAll the same principles apply that you just learned to a biological system. Answering the question of how B12 is reduced in the body./nB12 is reduced in the body by a protein called flavodoxin, and this contains another B vitamin, a flavin B vitamin./nIf we look at the standard cell potentials, the reduction potentials for these reactions, the reduction potential E knot for vitamin B12 is minus 0.526 volts. That is a very low number for a biological system./nThat is one of the lowest standard reduction potentials that is known./nFlavodoxin is minus 0.23 volts. Let's think about this reaction and whether it is favorable or not. B12 then is a better reducing agent than flavodoxin. It has a more negative number./nAs we just saw in the last example that makes it a better reducing agent./nIt is hard to reduce B12, easier to oxidize. Flavodoxin has a more positive number. Just like we saw in the last one, it is not as good a reducing agent. If this reaction occurred spontaneously, B12 should be reducing flavodoxin, not the other way around./nWe just worked an example where we saw there was one thing that was a really good reducing agent and another wasn't as good./nAnd we had an enormous K, so it was very much favored in the other direction./nThis should not be occurring. It doesn't seem possible that this is occurring, yet it is. Is the reduction spontaneous? No, it wouldn't be. Let's actually calculate the numbers and see how bad it is./nAgain, you can use this exact same equation. Instead of calling things cathodes and anodes, you can just consider it in terms of the reduction and the oxidation./nAnd we can plug in the values. What is being reduced is B12./nWhat is being oxidized is flavodoxin in the body. We can put in the numbers minus 0.526 volts. And then it is minus a negative number of minus 0.23 volts. And we get out that the cell potential is minus 0.296 volts./nThat is not going to be spontaneous./nIt is a negative number for delta E of the cell, which is going to mean what about delta G? It is going to be positive. No, it is definitely not spontaneous. And then we can continue to go on and plug in these numbers and calculate what delta G knot would be./nBy this equation, that we have seen many times today, it is a one electron reduction, so one electron is involved./nWe plug in Faraday's constant and we plug in the cell potential, and we get plus 28.6 kilojoules per mole./nThat is a very positive number. It is a pretty big number for delta G. So how does this reaction go? Why don't we all have heart disease, magaloblastic anemia, problems finding our way home?/nWhat happens? Well, the answer is that the body has to apply energy to get the reaction to go./nIn this case, the energy comes in the form of a molecule called s-adenosylmethionine. And the cleavage of s-adenosylmethionine generates the delta G for that is minus 37.6 kilojoules per mole, so it is very favorable./nAnd it is more favorable than this reaction is unfavorable./nWhen you put those things together it goes. What do you call it when you have a cell in which an unfavorable reaction is driven by applying energy? What was that called that we talked about last time? Yeah, electrolytic cell./nThe body just does the same thing that you would to drive any non-spontaneous reaction./nIt comes up with an energy source, couples the reaction and then the reaction goes. Again, same principles that apply.
Tags // Oxidation Reduction
Added: April 16, 2009, 11:32 pm
Runtime: 2890.07 | Views: 13349 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - The Shapes of Molecules: VSEPR Theory
Principles of Chemical Science/nVideo Lectures - Lecture 29/nTopics covered: /nThe Shapes of Molecules: VSEPR Theory/nInstructor: /nProf. Catherine Drennan/nTranscript - Lecture 29/nMonday's class really picks up from the lecture we had this past Monday. We started crystal field theory and then talked about octahedral-shaped molecules and splitting of d-orbitals. And then on Monday, we are going to continue with that topic./nWe are going to move in and talk about tetrahedral molecules, square planar molecules and how they are going to split the d-orbitals. And then how that relates to something, why we need to know about any of that. And that is because it is able to predict some of the properties of molecules, like their colors and also their properties in terms of spin and using various different spectroscopies to learn something about particular molecules, coordination complexes./nBefore we get into that unit, we are going to kind of take a step back and talk today about shapes of molecules. We introduced octahedral, but now we are going to talk about all the other kinds of shapes of molecules in today's lecture. This is kind of jumping around in the book a little bit. We have been in Chapter 16. And now, the way the book is divided, this particular unit is back in Chapter 3, although it is referred to a lot in Chapter 16. This is kind of some of the things that you need to know to finish up Chapter 16. And then on Monday we will be back in Chapter 16 as well. Again, this is a little bit odd order because of the way the calendar fell this year. All right. VSEPR, shapes of molecules, why do we need to know about this?/nWell, as most of you know, I do protein crystallography, which is all about structure. So, of course, I am going to think that structure is important, that shape and geometry are important. And, in fact, they have an influence on the physical properties and chemical properties of molecules. Things like melting points, boiling points, reactivities, and, in particular, in biological systems it is important./nA biological system has particular shapes of chiral molecules, it is a chiral environment, which we have talked about before. And so molecules must fit precisely in that environment. Often, if you are talking about an inhibitor of an enzyme, the inhibitor or the pharmaceutical drug is designed so it specifically fits into a particular crevice in an active site and blocks activity. A lot of the world of pharmaceuticals is all about shape of molecules and how you design particular types of shapes of molecules./nWe have learned a lot about this sort of area already, and we found that the shapes of molecules can be predicted pretty well using simple methods based on Lewis structures and also in terms of thinking about sort of where orbitals are and how that may affect the particular shape of the molecules. Again, this is sort of a simplistic view that works pretty well. And it will be also, on the next problem set, a good opportunity for you to review Lewis structures before you take the final. Sometimes this unit is with Lewis structures, there is some rationale for that, but it is also kind of nice to come back to it later in the semester as a way to review things that you learned before and get you ready for the final exam. The particular theory we are going to talk about today is referred to as VSEPR. And this is the valence-shell electron-pair repulsion theory./nAnd so, like crystal field theory, you are talking about repulsive effects. Here this is based on a simple principle that valence electron pairs will repel each other and that the geometry around a central atom will be such to minimize that repulsion. This is a lot like crystal field theory, where we were talking about the splitting of the d-orbitals. And we are talking about ligands as negative point charges heading toward those d-orbitals. A lot of these theories that work pretty well are just really based on the idea that repulsion is unfavorable. It is a pretty simple concept which can explain a lot in terms of properties of molecules. Before we get into this, there is always nomenclature that we must deal with. And here is the nomenclature here. We have A as a central atom./nWhatever is in the middle, which when we are talking about transition metals, complexes would be your metal, but this works for organic chemistry as well, when you are all talking about carbon, oxygen, sulfur, nitrogen. And X is going to be your bound atom, and E will be a lone pair of electrons, hence the E. There are just a few rules. It is not too bad compared to some of the other things we have done. Just a handful of rules that you need to know. First, you need to know that the steric number, abbreviated S.N., is used to predict geometries, and what steric number refers to. It refers to the number of atoms bound to central atom, plus the number of lone pair electrons on the central atom./nHere we are not concerned about double or triple bonds. We are only concerned about how many things are around our central atom, both the bonded and the lone pairs. Whether it is a double or a triple bond doesn't matter at all for these cases. It is the number of atoms bonded, not the number of bonds, which is important to remember. And so steric number, here, should not be confused with coordination number, which we talked about in transition metals, --/n-- which is just the number of ligands around the central metal. So, there are a number of different things that you are going to learn about. All right. For resonance structures, you can imply a VSEPR to any one of them. And there are two cases that we need to consider today. One is the case when there are no lone pairs, and other is the case when there are lone pairs. Those are all the rules. It is really not too complicated for this. We are just going to go through today and look at all these different molecules and consider the nomenclature and the geometry around them. First, we will consider formula type./nAnd so AX two would mean that you have a central atom A, and we have two atoms bound to it, so two Xs. And now we are considering the cases without lone pairs of electrons, so there won't be any Es until later. Then the steric number is going to be two. And so here would be the shape of the molecule. And so we have a center atom and two atoms bound to it. It would have linear geometry and a bond angle of 180./nThat is pretty simple. Now, we go to the next category, where you have three things bound, so that would have an SN number of three. And this is going to be the shape of that molecule. And what is this called? Trigonal planar. And what are the angles? 120. Okay. Now, if we move up to four, and this one should be pretty familiar to you, if we have four things bound, SN number of four, this is the molecular shape that we have. And you all know this as tetrahedral. And the angles of tetrahedral are 109.5. These are all things you have to memorize./nThis is written in your notes, but if you don't look at your notes and just try to think about it, it will be a good review for the next exam on this topic. All right. If we go into the next category, when we have five things bound, that would be AX five, five atoms bound, SN number of five. We have this trigonal bipyramidal geometry. And what are the angles around the equator here connecting the red atoms? 120. And what about red to black? 90. All right. Then, if we go to the next category, when we have six things, we talked a lot about this in the last two classes, what kind of geometry is this? It is octahedral geometry, with angles of 90./nRight. These are all the basic shapes when you just have two through six atoms bound. Let's take a look at a couple of examples of molecules when we have actual atoms listed. If we look at CO two, what is our SN number? Two, right. We are only considering the number of atoms bound. We don't care that they are double bonds. We are only concerned with the atoms that are linked to the central atom here. Then the geometry is what? Linear. And angles? 180, right. We will look at another example here. Is this one going to be the same or different from the one above? The same, right, because we don't care about the double or triple bonds. We can go through. SN number is still two./nIt is still linear. And the angle is still 180. We move into another example here. And I will just say that these lecture handouts have this information in here. This is the most requested lecture handout that I have from people who have been in the classes who get up to other level classes and need to know this stuff. And, for some reason, some of them have not kept their 5.111 notes sort of next to their pillow in later years./nI am not really sure why, but sometimes it gets lost and they request this handout, because it is a really nice summary of all of these things. You might want to keep it in a nice special place for future reference. All right. Some day I should bring in what my freshman chemistry textbook looked like with like everything highlighted and stuff just for entertainment value. All right. This molecule here would have a SN number of three. And its shape again? Trigonal planar. And angles? 180. By the time you get through with lecture today, you should have all of this memorized, because we are going to see these a couple more times. All right. This one, what molecule is this? Does someone know its name? Methane, right. Again, what is this geometry?/nTetrahedral. SN number of four. And angles of? 109.5, right. Okay. Now we get into a few more complicated molecules that you do not see quite as often, although I think we have seen this one quite often in different kinds of problems, including chemical equilibrium and other things. SN number here? Five. Geometry? It is hard to say that one in unison. We can work on this for the end of lecture. Then we can all say it in harmony. All right. And our angles again? Yeah. Okay. And one last one. This should be very familiar to you. This is SN number six. Geometry? Octahedral. Angles? 90. Okay. People are doing well./nHere is some stuff that you may not have heard before, or maybe have in another chemistry course, but now we are going to introduce lone pairs and see what effect this has. It gets a little bit more complicated at this point. For lone pairs, we think of them as being sort of centrally located. And I brought in some props to talk about the spatial distribution of electrons. One can imagine that the electrons are spread out as far as they can go. They like to have a lot of space. And so the lone pair electrons are going to take up more space and experience more repulsion then when you have the electrons as part of bonds to atoms. The lone pairs are not sort of held in as much, they are not sort of forced to be in one location, so they will spread out more./nAnd so because of this, their spreading out more, they are taking up room, they are kind of pushing everything away, because they want to have as much room as they possibly can. That means that lone pairs are going to be more repulsive than the atoms that are bound. Even though electrons are involved in those bonds, they are not quite as bad. The worst thing you can have is two sets of lone pairs near each other. They want all the room./nAnd so they are going to be sort of bumping into each other and trying to force the other set farther away. See, this is much more bulky than if you just have the two atoms bound with bonds to a central atom. The next would be a lone pair versus a bonding pair. And then the least repulsive will be two bonding pairs because those electrons are sort of fixed into locations for the bonds. They are not requiring as much space./nThey are not sort of spreading out and taking over all the space that is available. Worst is lone pair-lone pair, then lone pair bonding, and then bonding-bonding. And, if you think about this sort of simple rule, you can predict a lot about the shapes of molecules. Let's look at what shape something will have if it has this sort of set molecule. Again, A, the central atom. X is four atoms that are bonded to the central atom, and E is one set of lone pair. And so this is going to have what is known as a seesaw shape. And let's think about why that would be the case. If we have this sort of central framework, but one of these bonded atoms is now going to be replaced with a lone pair. There are five things around this framework./nOne is a lone pair. Where are you going to put it? If we put it on top, what are the angles between these lone pairs on top and the three red bonded atoms? 90. That is kind of close and not so favorable. If we instead put one over here, then we have still the sort of negative interaction with the 90 here and the 90 here. But we only have two of them instead of three. And these guys are 120 apart, so it is not so bad. There is a little bit more space in the equatorial position. And so that is why you get this seesaw shape, because you would want to put it in the equatorial position, where there is only sort of these two rather than three./nThere is just a little bit more space, because these guys are farther away. That gives rise to seesaw geometry. You can sort of again, these rules are more to explain in retrospect what is observed about those molecules, and then you rationalize. But the rationalization works pretty well. If we do that, if we put a lone pair, then, here, instead of one of these top positions, we get this seesaw geometry./nAnd so let me help you remember this. See? Get the seesaw geometry. How many of you actually grew up with seesaws in the neighborhood playground? Still most people, okay. They are actually one of the most dangerous toys. And I think a lot of playgrounds have removed them because of the one kid who sort of sits on it and keeps you, like, stuck up in the air because they are a little heavier than you are./nAnd you are stuck there for a really long time, and then eventually they get up and you go falling and flying. We may have to give this a new name at some point when all seesaws are removed from playgrounds, but it sounds like it is still usable. Now, if we wanted to put another lone pair on that same structure, we still have essential atom, three atoms bound. Now we have two lone pairs. Those two lone pairs, then, will go into the equatorial positions as well. And so you kind of have this sort of arrangement. And they were so repulsive they actually made the atom unbonded from it. Okay./nThat leaves you with this shape, which is a T-shape molecule. Again, this is a rationalization. Someone figured out that a molecule like this had this T-shape ad it could be rationalized by the fact that the lone pairs wanted to sort of fill out in that equatorial plane. And they could move away slightly from each other, and that would minimize the amounts of repulsion. All right. If we keep going here and look, now, at a molecule with this geometry, AX four E two molecule. That is based on kind of what structural framework? What is the sort of parent structure when you have six things bound? Yeah, based on an octahedral geometry./nIf you take your octahedral molecule and if you want to figure out where you are going to put your lone pairs, you could stick one down here in the equatorial, but that doesn't do you as much good this time because now your other equatorial things are at 90. This one would have a lot of crashes down here. What ends up happening is the two lone pairs would go one above and one below, so they would be as far away from each other as possible. You have one below and another one on top. And remember, the worst thing to have is a lone pair-lone pair. Even though you have four atoms at 90 that it is interacting with, at least it is far away from the other lone pair./nEach kind of gets half the molecule, one up top and one on the bottom, so they have the most room, those electrons, to spread out. And so that gives you what kind of geometry? Square planar, right. All right. You can also predict something about what might happen to the angles that you have all now pretty much memorized on the previous slides. That if you do have a lone pair, and out here we have an AX three E, SN number of four, so four things bound, one lone pair and three atoms, and this is based then on what kind of structure? With SN number of four, what is the kind of the parent geometry there? Yeah, tetrahedral is the parent geometry./nAnd now we have replaced one of the bonded atoms with a lone pair. And so what that is going to do is that lone pair is really not in that bad of shape. You put it on top. You have 109.5 now between those angles pointing down in the parent geometry. And that was better than the others which were at 90, but it is still not good enough for it. And so it actually presses down even more. Those lone pairs, they are going to spread out as much as they possibly can and repel those bonded atoms away to the extent possible. In this molecule, what is observed is that the angles in between here, in between the hydrogens, instead of being 109.5, which is what you would expect for the true tetrahedral geometry, they are 106.7. This lone pair is pressing down, compressing the bonded atoms even closer to each other than what is sort of the normal framework./nThis really takes up a lot of room here and is repulsive toward these bonded angles. Again, the VSEPR theory can predict these kinds of, or really explain these kinds of changes. If you observe the smaller angle you can say, oh, there was a lone pair. That makes sense that it would be a smaller angle because there would be repulsion down. Again, you can think about it. You can make one of these lone pairs yourself, low tech and think about the shapes here. All right. We can also rationalize the difference between a PH3 and NH3 molecule. They have pretty much the same framework here, based on tetrahedral geometry AX3E, SN number of four./nBut now, instead of the number for methane that we saw of 109.5, and instead of the 106.7 for NH3, we now get a really small angle between here, between the angle between hydrogen and phosphorus. And hydrogen is now 93.3, so it is even more compressed. And the reason why is because we are moving down in the Periodic Table. And so as we move down the column in the Periodic Table, we are going to have lone pairs that have even larger spatial volumes. And so they are going to compress even more. Now it is an even bigger sort of situation here. The lone pairs are trying to fill up even more volume./nAnd so they are sort of trying to compress as much as possible these atoms away. They are filling up all available space and making the angle here even smaller. And so if asked about this, if asked to predict, you could sort of look at where things are in the Periodic Table. Here is nitrogen, here is phosphorus. And so if we had a compound based on nitrogen or phosphorus that was based on tetrahedral geometry, one might expect that the angle becomes even smaller as you move down the Periodic Table, as that lone pair that is on there is filling up even more and more spatial volume. This kind of combines ideas of trends in the Periodic Table with shapes of molecules. All right. Another thing that you can sometimes predict is about bond lengths./nIf you experimentally observed different bond lengths in a molecule, you could use VSEPR to go back and think about where that may have come from. What is this geometry here, the one that is shown and also the parent geometry? The one that is shown is what? Seesaw. And the parent geometry? Right, trigonal bipyramidal. All right. Here we have AX four E. We have an SN number of five. We have four atoms bound. We have one lone pair. And now it is observed that there are two sets of SF bonds. One set is 1.545, and one is 1.646. What do you think those different bond sets --/nThere are two types of bonds. We have one lone pair. Likely one set would involve the gray atoms, and the other would be the black atoms. Which do you think would be longer? Do you think the gray or the black would be longer, the bond lengths from the central atom to the gray or to the black would be longer? Gray. And why is that? Right, because these are at 90 going down. Whereas, this is at 120. The 90 is more repulsive, so it basically kind of pushes that up farther and down farther, moves them farther away. The black atoms are already farther and won't feel the repulsion as much./nAnother way to sort of, if you have a lone pair here, for the lone pair to feel like it really has the space that it requires, is to sort of repel the atoms that are closer in terms of distance, repel them just a little bit farther away. And then it can kind of fill out and feel like it has enough space. Again, if you had an experimentally observed set of bond lengths, you could rationalize using VSEPR, why you would have this./nOkay. As we said, more repulsion between the lone pair electrons and the axial ones because those are at 90, so those would move farther away. That is a lot of the theory. Now we are just going to run through these sheets and talk about the different types of geometries, the names of the different types of geometries when you have the lone pairs on there. And then look at some examples. Again, if you call out the answers, and if we go through this quickly, this is all that we need to go over for today, and it will be a good review. You will sort of have the opportunity, today, to memorize this. You might want to go back over it a couple times, but pretty much this is the whole unit. It is not too challenging. If we have AX two E, we are going to have a SN number of three./nAnd a molecular shape. I will draw that. And what is the name of that geometry? It is bent. And that is because you are only considering the bonded atoms. You are not talking about the lone pair. We don't really see the lone pair, two bends, and so it would just look bent to us. That is where that comes from. And what angles would this have for this particular case, SN number of three? Okay. We have this one here. The SN three would have a trigonal planar parent geometry./nBut if we take a lone pair and put that on instead of an atom, then it will also be somewhat compressing of those two atoms down there. We won't really know, without knowing what the central atom is, we cannot really predict how bad the compression is going to be. What you would put is less than 120 here, because it would be 120 if this was an atom and not a lone pair. If this is a lone pair, then it is less than 120./nAll right. Here is another case. Now we have three atoms bound, one lone pair, SN number of four. And our parent geometry here would be tetrahedral. And now we have a lone pair. We saw this a little bit before. And what is that geometry called? Here, we have this picture. What is this called? Trigonal pyramidal. We have this. And what would the angles be here? Less than 109.5. All of you should know the answers because they are actually in your handout, so feel free and confident to yell them out. All right. Again, this is going to be less than 109.5 because we have a lone pair pressing down./nYou have to in all these cases consider what the parent geometry is. And then the angles are going to be a little less than that if you have one lone pair sitting on top. All right. AX two E two has a SN number of four. All right. What geometry is this based on? Yeah, tetrahedral again. We have two lone pairs on that. And what does that look like it should be called? Bent as well. And now the angles are going to be less than 109.5. So, you see, we have two bent molecules, but the angle isn't the same. It depends on whether it is based on a trigonal planar or it is based on tetrahedral. You need to sort of know what the whole molecule was to figure out what degree of bend you have in these two./nAll right. AX four E has SN number of five. And so our shape here, what is this one going to be? Yeah, this is going to be a seesaw. And so what are our angles going to be? Right, less than 120, less than 90. It could be that really the 120 are not going to be that affected, but that will be what you will put down because you don't really know. This is just a rough predictor of what those are going to be, so it is less than what the parent molecule was. All right. SN number of five. And so what are we going to be doing here? We are going to add another lone, take off an atom, add another lone pair./nAnd what are we going to get here? Right, T-shaped. And the angles will be... At this point is our T-shape, and so we will have less than 90. You will have the two lone pairs in the equatorial position, as we talked about before, trying to spread out as much as they possibly can. Now we are replacing another one with a lone pair. Another bonded atom. Our SN number is still five. And what shape are we going to get in that case? Linear, right. They all go in these equatorial positions to spread out as much as they possibly can. And so we are going to end up with a linear molecule, so we would replace here. And what are our angles here? They are exactly 180. Yeah?/nIt could be that there would, but they probably will just be really spreading out around to fill out as much room as possible around the whole equatorial place. You can just sort of think about it as if there was space, they would move into it, so they have really filled up this space. And pretty much you might as well keep them there, because whichever way they move, they are just going to get repelled in another direction./nThat's kind of the one exception, are these ones that turn out to be linear. There is one other exception where the angles are not going to change as well, where you have orbitals on kind of either side that counteract. That is a good question. All right. One more. We have AX five E, SN number of six. This is based on what parent molecule? Octahedral. We are going to take off one ligand, and so this would be, if we removed one of the ligands here, we get this parent shape, which is called square pyramidal. And what are the angles going to be, here? Less than 90. Okay./nEvery year, when I do this unit, I like to reward people who are in class. And so I give a little hint about the final. And the hint will be that this will be on the final, [LAUGHTER] so you might want to put that down. The last couple of years I did seesaw, which really is my personal favorite, but at some point I have to switch. This time it will be this one. And it is always so sad to me when people get that one wrong, so please get that one right. Okay. Let's keep going. SN number is six. And so we are going to put on another lone pair. And where are we going to put it? We are going to try to move it so that there is the least amount of repulsion between the lone pairs, so it is going to be up on top./nThis geometry keeps the lone pairs as far away from each other as possible. And what is the remaining thing called when you have those lone pairs on there? That is going to be square planar. And what about the angles in this case? In this case they are just 90, because you have a lone pair on top pushing down and a lone pair on the bottom pushing up. This is another example where it is not less than. All right. Next one. We add another lone pair. And at this point does it matter where you put the lone pair? No. They are all equivalent positions, so it doesn't make any difference. It is going to be one of these guys gets replaced, and so that is going to leave you with the what geometry?/nT-shaped. And angle here is going to be this time less than 90, right, because it will repel this one and this one if you have a big lone pair in this position. All right. One more. Again, the same geometry. We take another bonded atom off and put another lone pair. And what are we left with in this case? Linear. Again, this is going to be 180. All right. Now we have looked at all of those, and we can very quickly go down. And in this handout, I have most of it written. I don't have the angles, so that you can write the angles in and have your complete set for this as well. Now let's look at water. The Lewis structure for water, you would end up with two lone pairs on the oxygen./nAnd so what formula type would this be? AX two E two. SN number of four. Geometry? Bent angle. Less than 109.5, because this is based on the tetrahedral geometry. Remember, bent could either be less than 120 or less than 109.5, but if you have a SN number of four then it is based on tetrahedral geometry. It would be less than 109.5, which is the tetrahedral angles. Okay. SF four, here is our structure. Formula type for this would be? AX four E. And that is a SN number of five. And the shape? Seesaw./nAnd angles? Yeah, less than 90 and less than 120. Here is another one, BrF three. These are all probably ones that you did Lewis structure for. If not, I am sure you will before the final. Here is another one. The formula type for this would be AX three E two. And SN number? Five. And geometry? T-shaped. Angles? Less than 90. Okay. If we keep going, here is another example, selenium and fluoride. And we have three sets of lone pairs. What is the formula type? AX two E three. SN number of five. Geometry? Linear. Angle?/n180. Okay. BrF five. Formula type? AX five E. SN number? Six. Geometry? Square pyramidal. Angles? Less than 90. You want to remember that one. Okay. Here is another one with an addition lone pair. Formula type? AX four E two. SN number? Six. Geometry? Square planar. Angle? 90, exactly. That completes that list. Again, the theory of VSEPR, you are taking experimental results and then trying to apply a theory to it. It is not good at really making exact predictions about things./nBut it is really good at rationalizing what they have. All right. Have a wonderful Thanksgiving, everybody.
Tags // VSEPR Theory
Added: April 16, 2009, 11:33 pm
Runtime: 2598.40 | Views: 25215 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - Transition Metals - Lecture 30
Principles of Chemical Science/nVideo Lectures - Lecture 30/nTopics covered: /nTransition Metals/nInstructor: /nProf. Catherine Drennan/nTranscript - Lecture 30/nYou all should have these handouts ready to go now./nLet's review where we were last week. If everyone can settle down a little bit, you can discuss about who wants to be the top chemistry major later on after class is over./nAll right. Octahedral case./nHere is our octahedral molecule. And, again, crystal field theory is just an explanation for the things that are observed experimentally. It is a fairly simple theory. And, if you go on to study chemistry, and you will take more inorganic chemistry later on, you will learn about more complicated theories./nBut this is really simple./nThis simple theory just thinks about each ligand, each small molecule like NH3 group bound to a metal. The metal is in the center position. That each ligand considers as a negative point charge. And so then the theory looks at whether that ligand, as a negative point charge, is pointing toward one of the d-orbitals./nYou need to know about the shapes of the d-orbitals./nAnd you need to know about the shapes of the coordination complexes. And then you need to just ask yourself, is that ligand pointing directly toward a d-orbital./nIf it is then it will be destabilized. If it is not then those orbitals will be stabilized. And so you are talking about the stabilization relative to an imaginary spherical and crystal field where ligands are just sort of everywhere all around the center metal./nYou consider that as sort of your standard case where all of the d-orbitals have the same energy./nBut then when you consider those ligands as negative point charges in particular locations, some d-orbitals then get destabilized and other ones get stabilized in terms of energy. In the octahedral case then, and here is our little cartoon for the octahedral case where there are ligands along the z-axis, ligands along the x-axis and ligands along the y-axis./nAnd so if you look at dx2 minus y2, you see that the ligands will be pointing directly toward these d-orbitals, because ligands are coming along y and along x./nAnd so there will be a lot of destabilization. For dz2, you have ligands pointing directly toward those orbitals as well along the z-axis. These have large repulsions due to the negative point charges./nThose negative point charges are heading right toward those d-orbitals./nAnd these two are degenerate with respect to each other. They have the same energy, the same amount of destabilization. And these are destabilized compared to this other set of d-orbitals./nAnd in this set of d-orbitals, none of the orbitals are on axis. They are 45 degrees off axis. The negative point charges are on axis. There is not as much destabilization. The ligands are not pointing directly toward any of these d-orbitals./nYou have smaller repulsion./nAll of these three of the same energy, they are degenerate. And they are all stabilized compared to this set. That is what we talked about last week. And we also talked about how you can draw an octahedral crystal field splitting diagram./nAnd here is the average energy of the d-orbitals with this sort of spherical crystal field where ligands are everywhere./nAnd so when ligands are everywhere, all of the d-orbitals have the same energy./nBut then when you put the ligands in their locations, in the octahedral geometry framework then these two sets of orbitals, dx2 minus y2 and dz2, are up in energy, they are destabilized. And these three go down in energy./nAnd the overall energy of the system is maintained./nHere is the crystal field splitting energy for the octahedral case, O for octahedral. And so it is split. Two orbitals go up in energy, so they go up by three-fifths, three go down, so they go down by two-fifths to maintain the overall energy of this system./nThat is what we talked about last time, and I just wanted to review it because we are going to go on and talk about other types of geometry. And, of course, remember that Exam 3 is on Wednesday. Extra problems are posted./nI will mention that I actually posted the original exam last year, and the exam that people took was slightly easier than the one I posted./nI did that actually by mistake. But it turns out that if you do the exam that is posted it is a better indicator of how well you will do on this next exam. The third exam last year was a bit too easy, i.e., too easy meaning the sort of outstanding category was like 98% and above./nAnd if you got a 97% then that was no longer outstanding, so whether you got three significant figures wrong or two made some difference in your grade./nThat clustering I don't like. I like to have the average such that people's performance on the exam really indicates more how well they know the material. The exam that is posted is actually the better view./nEven though that was an accident, I will pretend I did it on purpose because it is a better thing. So, extra problems on./nToday, we are going to finish up crystal field theory and then talk about the relevance of crystal field theory./nAnd the relevance comes in to explaining the colors of these transition compounds. Compounds with metals have some really spectacular colors. And you can explain and rationalize some of these colors by this really very simple theory, so that is fun./nAnd you can also explain some of the properties and magnetism of these compounds by this fairly simple theory./nOK. Let's look at the tetrahedral case now. Here we have our tetrahedral molecule. I will try to kind of hold it up as it is displayed on the board./nIn this case, we have two ligands up, one coming towards you, one going back. And then we have two ligands in the plane of the screen./nAnd, in this case, the convention is to draw the z-axis coming down here and the y-axis is kind of coming in from the side and the x-axis is coming out toward you here./nThat is what this picture indicates. The crystal field splitting energy tends to be smaller than the octahedral crystal field energy because none of the ligand point charges are really pointing toward any of the d-orbitals./nLet's just take a look at all the d-orbitals and consider then which ones are going to have more repulsion than other ones./nIf we go through and take a look at it, and I have tried to indicate, here are our tetrahedral cases and here are five d-orbitals. And these little lines, the negative charge and the arrow is sort of a rough approximation of how those ligand point charges are going to be coming in and pointing toward the d-orbitals if they are coming in with this tetrahedral geometry as drawn./nAnd so you can see for the dx2 minus y2, none of those ligands are pointing right any of the d-orbitals, and they are definitely not for the dz2 either./nA little bit more here but, again, not exactly pointing toward them. Although, this picture looks like it is a little bit./nAgain, it is closer here but not exact. The result of this is that there is more destabilization down here than there is up here./nIt turns out that this is then opposite of the octahedral case that we just talked about. And octahedral, these are the ones that are most destabilized. In tetrahedral these are stabilized compared to the others./nIt is flipped exactly around./nAgain, there would be more repulsion between the ligands as negative point charges and the orbitals that are 45 degree off axis. Because here the ligands are off axis and also the orbitals are off axis./nThey are not directly pointing at each other but it is definitely worse down here. These guys are really far away from those ligand point charges./nThese two will have the same energy with respect to each other./nThey will also be degenerate as they were before. And these three have the same energy with respect to each other. They are also degenerate. Then we can draw what the diagram will look like for the tetrahedral case./nAgain, here was the octahedral case. Now, this is the tetrahedral case./nAnd so we have the splitting energy, which is the energy that splits the orbitals based on these negative point charges compared to this theoretical spherical crystal field./nAnd we also have a tetrahedral crystal field splitting energy. That is drawn here. Now, the first thing, which I emphasize that I will point out again, is that these guys are flipped around./nInstead of being destabilized as they are in the octahedral crystal field, the dx2 minus y2 and d dz2 are now stabilized compared to the other orbital set./nThey have switched positions. These, instead of being stabilized, are more destabilized. It is a direct switch. There is also a little change in nomenclature here. You can still refer to this orbital set with the term E./nBut instead of Eg we have E./nAnd I will just point out that the book calls this E2 and sometimes it calls it E. It should just be E. This is kind of a small point, but I don't want people to get confused by that. And instead of t2g, we have t2 up here./nBasically, the g term is gone in this case. And you will learn a lot more about that if you go on and take more inorganic chemistry. These are flipped./nWhat else is true? Well, the other thing that is true is that this splitting energy is much smaller in the tetrahedral case than it is in the octahedral case./nHere there is a much bigger distance between the orbitals. Here it is a much smaller distance. And that is because in the octahedral case there are ligands pointing directly at the d-orbitals. In this framework for octahedral there are ligands right along z, along x and along y./nAnd so they are along all of the orbitals, pointing directly toward the orbitals./nIn the tetrahedral case they are not really pointing directly toward. They are closer to some than to others, but it is not as direct. A result of that is that this is just not split that much. There is not such a huge difference between the energy levels of the d-orbitals in this case./nWhat does this mean?/nWell, it means that a lot of tetrahedral complexes, because the splitting energy is small, are called high spin. And we talked about this last Monday that in high spin you have the maximum number of unpaired electrons./nAnd because of the purpose of this you can just assume that if you have a tetrahedral complex is will be high spin./nThat will come into play when you figure out how to put your electrons into this diagram, whether you put them in singly to the fullest extent possible without pairing or if you start to pair./nIn this case, you can put them in singly to the fullest extent possible. OK. As I mentioned before, the overall energy of this system is maintained. In this case, you have two orbitals that go up in energy./nThey go up by three-fifths./nThree that come down, so they come down by two-fifths. In this case, you have three going up so they will go up by two-fifths. And you have two going down in energy, so they will go down by three-fifths to maintain the overall energy of the system./nThat's the tetrahedral case./nNow let's consider square planar./nHere we have our square planar molecule and we have ligands along the y-axis, ligand along the x-axis pointing out towards you, one, and one going back behind. And z then is coming up and down in this plane./nWith these ligands along y and along x, which orbitals do you predict would be destabilized the most? I heard dxy. That is one. What else? dx2 minus y2, right. Those two. The ones that don't have the z component in them./nThere is really not much destabilization along z. There are no ligands along z./nOrbitals that have amplitude along z are not as affected. The ones that will be most affected are the orbitals that have component in the x and the y direction where the ligands are./nNow we can consider the d-orbitals and talk a little bit more about this. If we look at these two sets of orbitals now, here, as you correctly predicted, there will be lots of repulsion./nThe ligands are pointing directly at these d-orbitals./nThe ligands are on axis. The orbitals are on axis. It is a direct hit. There is a lot of destabilization. And this one now is going to be destabilized compared to all other ones. It is going to be by itself, the most destabilized orbital./ndx2 doesn't really have much destabilization at all./nIn fact, it will be stabilized. There are no ligands along z, there is no real repulsion of the ligands to dz2, so it is going to be a lot less than for this guy and also less than dxy./nWe can draw this now, put in two of the orbitals. This one is going to be way up. It is going to be by far the most destabilized./nIt is going to even go up from where it was in the octahedral case./nWhereas, dz2, which was up here, degenerate with dx2 minus y2 in the octahedral case is now going to drop way down. These are not equivalent at all anymore. And so this makes sense. The difference between the octahedral geometry is that there are ligands along z./nNow the ligands along z are gone in the square planar case, so there is a big difference between these two orbitals sets in terms of energy./nNow let's look at these three. This one, as people correctly predicted, is going to have a lot more repulsion than either of these because the orbitals have maximum amplitude along x and y./nThey are actually 45 degrees off x and y, but still we don't have anything in the z component here./nThis will be destabilized compared to these two. And these two are stabilized compared to the d-orbitals dxy, dx2 minus y2./nNow we can go and take a look at where these would be positioned approximately. If we look at this one, it is not nearly as destabilized as this one. Still, it is a bit destabilized and above this./nThis one is lower in energy./nAnd if you consider the difference here again, dx2 minus y2 the ligands are pointing directly at the orbitals. It is on axis. Here it is 45 degrees off axis, so it is not as destabilized but still is up above where it was in the octahedral case./nAnd then, finally, these two guys are going to be stabilized compared to this theoretical spherical crystal field. They are going to be somewhere down here./nThe big points here then are that, compared to the octahedral case, when you get rid of the ligands along z, this guys shoots up, this one comes down, it is stabilized./nThe other d-orbital that has maximum components along x and y, which is where those ligands are, goes up a little. These three are all sort of stabilized, these two are destabilized, but this one is destabilized quite a bit more./nThese also will have, you can draw sort of splitting energies here, but that is just too complicated./nWe are only going to do this with the octahedral and tetrahedral case because there are many more layers here. And these are little bit different. All you need to know is sort of this relative ratio./nThis is the most destabilized. This is next. These guys are stabilized. That is all you need to know./nWe are not going to draw splitting energies going up by one-fifth and things like that for this case, but you will hear about this more if you go on in chemistry./nWhat I expect you to know is to be able to rationalize. If you know something about the shape of a molecule, suppose you had a linear molecule along the z-axis, a good test question might be which orbitals would be the most destabilized, if you are linear, along z, for example./nYou should be able to think about the shapes of the d-orbitals, think about the shapes of the molecules and predict the orbitals with the most stabilization and the most destabilization and to be able to explain your rationale for predicting these splitting diagrams./nNow we are going to talk about how these diagrams and thinking about the splitting of the d-orbitals can explain the beautiful colors of these coordination complexes./nLet's go back to the octahedral case for a minute and talk about two different octahedral complexes./nWe have iron plus three in two different complexes. We have a high spin state with iron and six water ligands and a low spin molecule coordination compound with iron and six cyanide ligands./nThe first thing you are going to do, if you are going to be talking about the splitting, is to figure out how many electrons you are talking about or come up with the d-electron count./nLet's do that. If we have iron with six water ligands./nThe overall charge on that is three plus. What is going to be our, and this is on the board, too, but what will be the oxidation number of iron? Plus three./nThis is zero. And so it all has to equal plus three. Then the d-count is going to be equal to --/nAnd I will put up a Periodic Table. The d-count will be equal to what? Five, right. We look up the group number, which is eight, minus three is going to be five./nAnd our other iron complex will be the same. This one has a charge of minus three. Cyanide is what?/nMinus one. Iron would also have to be plus three to get the correct overall charge. Our d-count over here as well is going to be d5./nWe have the same number of electrons that we need to put into two different diagrams./nWe have one. This is our high spin case. And so we can draw the crystal field splitting diagram for the high spin --/n-- and put in our orbitals./ndx2 minus y2, dz2./nWe have our octahedral crystal field splitting energy. We are going up by three-fifths and down by two-fifths for these compared to this theoretical crystal field. And then we have our low spin case./nIn our low spin case, we have a larger splitting energy./nAgain, some are going to go up by three-fifths and three will go down by two-fifths./nNow we need to put in our five electrons./nHow am I going to put them in here? Will I fill up all of the different energy levels with one before I pair or will I start to pair in the lower first? We are going to put them in everywhere first./nWe will put all five electrons. We will put one in each of the orbitals./nI am going to place them singly in. That is because in this case, the splitting energy is small. And, since it is small, it is less than the pairing energy./nThe pairing energy is the energy required to pair up these electrons. There is a certain amount of repulsion. All of the electrons would love to be alone in their orbital. They would prefer not for you to put another electron in./nThere is a certain amount of energy required to do this./nIf the energy level between the d-orbitals is small then you put them all in singly and don't pair until you absolutely have to. This is referred to as the high spin case where you have the most maximum number of unpaired electrons./nIn this case, we have a large energy. That energy then is greater than the energy required to pair./nPE or the pairing energy. Now we are going to put them all into the bottom until we have to go to the top./nWe will first start out putting them in singly, one, two, three, and then we will start pairing, four, five. Now we fill up this bottom set. I should put in the t2g and the Eg labels. We will put them in this bottom set in t2g until we have to go up to Eg./nBecause this is a big energy difference./nAnd you don't want to do that until you have to./nAll right. Now we can look at the d-electron count, so where the electrons are. Here it is t2g3. There are three in the bottom./nAnd in Eg there are two./nAnd over here, in t2g, we have five. There are none in Eg. Those are our d-electron counts. And then we can calculate the crystal field stabilization energy, which shouldn't be confused with the splitting energy./nThis considers how much stabilization you have as a result of the fact that the d-orbitals are split in energy./nIn this case, you want to count up how many are in the lower energy levels. There are three down here./nThree times minus two-fifths the octahedral crystal field splitting energy. And you have two above, so two are up in energy by three-fifths. And so what does that equal? Zero./nAnd over here then you have five that are down in energy./nYou would have minus ten-fifths, the octahedral crystal field splitting energy. And sometimes in some books it also includes, there is some energy associated with pairing, so occasionally you will see this written as 2PE indicating that there are two sets with paired electrons./nThere is some energy associated with that./nIt is not as fantastic a situation as it look like. That there is lots of stabilization where the other has a zero stabilization, zero destabilization, it is just about the same, but there is some energy associated with this so sometimes you will see that./nAnd you only need to indicate that if it is asked for. And I think your book doesn't ask for it, but just so you are aware./nAll right. That is kind of a review. We talked about these things last Monday, but I just wanted to go through it again because now you need to see how this relates to colors./nWe have to introduce this idea that the relative ability of certain common ligands to split the d-orbital level is what is meant by this spectra chemical series./nWhich ligands there are, in one case, water and cyanide causes a difference./nAnd a strong field ligand, and we talked about this a bit last Monday as well, produces large energy separation. Over here we have a large energy, so this would be a strong field over here. It is strong./nIt splits those d-orbitals a lot. And a weak field ligand produces a small separation./nIt is pretty weak. It cannot really split those d-orbitals very much. It is weak. This would be a weak field case over here./nThere is not much splitting between the d-orbitals. Here are some various ligands and how they fall into this series. And you need to know this for the test, only these sets of ligands. There are three in the sort of weak field category./nThree in the intermediate category./nThree in the strong field category. When it is a weak field ligand, you are going to have the small splitting, as shown here. Therefore, you will have a high spin system where you have the maximum number of unpaired electrons./nIf you have a strong field, you have a large splitting. And then you are going to have a low spin system where you have the minimum number of unpaired electrons./nIn the example that we just had, this is high spin, so it is actually kind of an intermediate but it was determined to be high spin experimentally./nThe intermediate ones are a little harder to predict. These will be strong field. These will be weak field. These are harder to predict. Cyanide, according to this list, is the strongest field ligand./nThat makes sense that that was then a low spin system./nColors are going to depend on what photons are absorbed and the energy level it takes to get to that excited high energy state. Here it is a very small difference. Here it is a big difference. You would need to absorb differently and then also transmit differently to excited those electrons./nThat is going to give rise to the different colors./nLet's just take a look at this. Now we are reviewing back to the first half of the semester where you saw this. We have the energy of the absorbed light is equal to Planck's constant times the frequency./nAnd so we are back here again. We are talking about splitting energies between the d-orbitals, the amount of energy to excite and electron from one set of orbitals to the higher set of orbitals./nWe are back here./nBut now we are going to have a little bit of difference in that we are going to add another equal sign and say that the energy then is going to be, in this case with a transition metal complexes, octahedral coordination complexes, it is equal to this octahedral crystal field splitting energy./nAgain, we are going to be talking about exciting electrons here up to here./nAnd so that depends then on what the energy difference is between these sets of orbitals. We have seen this before, and now we are just adding this new term on the other end, the octahedral crystal field splitting energy./nAll right. This we also saw in the first half of the course. If you have a high frequency of light absorbed, the wavelength of the absorbed light is going to be short./nAnd that is that equation. I think these two equations were equations you need to have memorized./nThey were not on the equation C. That indicates that they are of really fundamental important. And then the color of the transmitted light is complementary to the color of the absorbed light. Here we need to think beyond the first half of the semester back to probably kindergarten./nWhen did you first learn about complementary colors? Second grade? I actually was talked into putting something like this on the exam so that you wouldn't have to have that material memorized./nAnyway, here are your complementary colors. If you know something about the color of the transition complex you can go back and calculate and estimate./nIf you are given actual numbers, you can calculate what the splitting energy difference was, or you can approximate whether it is going to be small or large depending on the colors that the complexes have./nLet's take a look at some examples here of two different coordination complexes, both octahedral. First we have to figure out the oxidation number./nAgain, this goes back to the last unit on oxidation-reduction, useful for the exam on Wednesday./nThe number here, you are going to have plus three for both of these guys. This is zero and that is zero. The overall charge is plus three, so chromium has to be plus three in both cases. Then when we do the d-electron count, we have to find chromium on our Periodic Table./nLook up what group it is in to calculate the d-count./nWe have that here. And we have a d3 case. What does CN stand for? Coordination number. And so what is the coordination number in this case? Six, right. There are six things here, six things coordinated./nAnd that is not to be confused with a steric number, which we talked about last Wednesday./nThe type of ligand here, what was what? What type of ligand in terms of weak to strong field? It was intermediate./nAnd over here? It is stronger. These guys are actually not that far apart from each other. Water was sort of intermediate. It is at least weaker than NH3./nNH3 was in the strong category, so it is at least stronger than water./nThey are not that far apart from each other, but this one should be a stronger field ligand than this one. If we consider a weaker field system and a stronger field system, is this going to make any difference to how we put in electrons this time? No, it won't because we only have three./nWe would put them in the bottom regardless./nThe issue is when you get to the fourth one. When you have a fourth electron you have to decide if you are going to put it up here or if you are going to pair, but there are only three so the diagrams look the same./nEven though the diagrams look the same these compounds are going to have different properties because they have different splitting energies./nAnd so the amount of energy required to pump an electron up here is going to be different than up here./nAnd so you will have different colors as a result. Let's take a look at that. Again, this is intermediate so this is smaller. This is stronger, it is large, it doesn't affect how the electrons go in, but it does affect the property of the compound./nLet's go through this and see what this means./nWater then is weaker. It means the splitting energy is smaller. If the splitting energy is smaller that means you have a lower energy, a smaller energy between them and a lower frequency that will be absorbed./nIn this case, this is a strong field ligand, so you will have a larger energy and a higher frequency that is absorbed./nOver here, if we have the lower frequency absorbed, we have a longer wavelength absorbed./nFor the stronger field case, we have somewhat of a higher frequency absorbed, so we will have somewhat of a shorter wavelength absorbed. And then the color of the transmitted light is complimentary to the color of the absorbed light./nHere you are going to have a transmitted light with a shorter wavelength./nYou had a longer wavelength absorbed so you will have a shorter wavelength transmitted. And this compound is, in fact, violet./nOver here you are going to have the shorter wavelength absorbed, so the transmitted light will be complimentary to that and we will have a longer wavelength. And this compound is, in fact, yellow. If we look again at the colors and their wavelengths --/nThis complex with the strong field ligand is yellow and the one with an intermediate field ligand is violet, somewhere down in here./nEven though water and NH3 are close to each other, the one that is stronger definitely has quite a distinct color from the one that is a bit weaker. And, unless you are given numbers, you cannot predict exactly which color./nYou can just rationalize the relationship./nIf you were told this is yellow and this is violet, tell me why, you would say this is a stronger field ligand than this one. That is the kind of thing you can do./nBut if you are given numbers and on problems, and on the final you may be given numbers, you can actually come up and calculate back and forth from the splitting energy to the color that you predict, from a color back to the splitting energy./nWhat things would be colorless?/nWhich kind of coordination complexes would be colorless, and why would it be colorless?/nThis is where things written in your handout can be read back to me. [LAUGHTER] Unless you know, but it is good to think about it a little bit./nOK./nIf you filled up all your d-orbitals and there is no room for things to move around, when it is not possible to have a d-to-d transition, you filled up everything so you are not going to be able to do this transition from one set of d-orbitals to another one then you have colorless./nHere are some examples of things that will be colorless, and let's just prove that that would be true, that those are, in fact, filled. We have zinc and cadmium./nIf we look, zinc and cadmium are in Group 12./nThey both have oxidation numbers of plus two, so twelve minus two is ten. You have a d10 system. This one is also going to be a d10 system. You are going to have all ten electrons already in. And so you are not going to have these transitions./nAnd so these, in fact, are colorless. Zinc is a really important metal in biology./nAnd it plays structural roles. It can play catalytic roles. It is a very important metal. And often, when biochemists are trying to figure out how a protein works, they won't realize it is a zinc containing enzyme because the protein won't have any color./nIn contrast, there are proteins and cofactors in metals that have a lot of color. And then, actually, as you are purifying the enzyme, it will have this bright orange or red color./nAnd you realize it has some metal cofactor in it right from the very beginning./nJust to go through then and sort of summarize this whole procedure, if you are on this end with the ligands in the weak field ligands then you are going to have a small splitting. Weak things cannot split much./nThey do not have enough strength to really split those d-orbitals apart./nYou will have a high spin system in this case, a maximum number of unpaired electrons. In terms of the complexes then they will absorb low energy photons, low frequency, long wave length, and then they will transmit high frequency or short wavelength./nThey are going to be in this violet blue-green end of the spectra. And let me just put this one up. Actually, I will put all these guys up./nThey will be on this end of the spectra. Again, it is all based on these two equations./nIf you have a small splitting small energy, you are going to have a low frequency. And then you will have a long wavelength for absorbed. And the complimentary for the transmitted. If you are a strong field, you will have a large splitting energy./nIt will be low spin, minimum number of unpaired electrons, and you will have a situation like this./nHere you fill them up to maximum extent. Here they are low spin, maximum paired, so then you will have the complimentary./nYou have short wave length absorbed. You will have long wavelength transmitted. And you will be on this end of the spectra, so you would expect yellow, orange or red in this case. You can make predictions./nIf you know something about the ligand, you can predict the color of the coordination complex, or at least rationalize the color of the coordination complex./nJust to show you sort of a real example, these are all octahedral complexes with cobalt./nAnd you can see the colors are really dramatically different. Depending on the ligand to the cobalt, you can really get this beautiful range of colors associated. And one of the courses that is offered during IAP that Dr./nShrank teaches, you go through and actually do some of these experiments and see these colors./nAnd so it is a little lab component that compliments nicely this part of the course, as well as a few other parts of the course./nThose are these coordination complexes. And cobalt is also in a vitamin. Can someone tell me what vitamin cobalt is in the center of? What is my favorite vitamin? B12, yes. These are colors of B12./nThis is protein bound to different forms of B12, so there are actually different ligands to the cobalt in these three cases./nAnd these are the actual colors. It is really spectacular, the colors of these. And these are crystals of the proteins. The colors are really very rich. When you are doing experiments on a B12 system, you know something about the form of B12./nIf you have done some damage to your B12, you know it by actually looking at the color of the molecular you are working with./nOr, the crystals or the protein you are working with. Very fast then, magnetism./nCompounds that have unpaired electrons are paramagnetic, attracted by a magnetic field, and ones where they are paired are diamagnetic. They are repelled by a magnetic field./nYou can figure out something about the geometries often of your compounds by figuring out whether the compound is paramagnetic or diamagnetic./nWell, how would you do this? Well, if, for example, you had a nickel two plus system in an enzyme, so you have a d8 center in your enzyme and it is found to be diamagnetic./nYou could ask the question, can I guess then whether it has square planar geometry, tetrahedral geometry or octahedral geometry? Because it is diamagnetic./nCan I figure out the geometry knowing that fact, knowing about unpaired or paired electrons? In this case, we would have to have its diamagnetic. All the electrons have to be paired. Here are our splitting diagrams for all three cases./nLet's look at octahedral./nAgain, we have to put eight electrons in. Here are eight electrons. Does it matter if this was weak field or strong field? In this case, which is it? Can you tell? Same, right, because you would put in electrons./nWhether you put them in one, two, three, four, five, six, seven, eight, or put them in one, two, three, four, five, six, seven eight, it is the same. This could be either./nNotice we have unpaired electrons, so this cannot be the geometry of this unknown center./nWhat about square planar? Let's put them in starting at the bottom. We will put them all in. Keep going and just put them all in. There are eight. It didn't really matter too much the order I put them in down here./nThe important point is, I really didn't want to put any in here until I had to because this one, if you recall, is really destabilized compared to the rest./nAll right. These are all paired up. And let's look at tetrahedral./nTetrahedral tends to be what kind of spin? High spin because the splitting energy is small, because you don't have ligands pointing directly at the d-orbitals. We can put them in, in a high spin fashion until we are done./nAnd that is our system. This is small splitting energy down here, so we have unpaired electrons here./nThis is paramagnetic, this is paramagnetic and this is diamagnetic. Then you could predict that the nickel and the enzyme must be square planar./nIt cannot be this and it cannot be that, so it tells you something about the geometry. And I will tell you that there are two square planar systems, at least, in enzymes of nickel. And people predicted it using similar kinds of techniques to what you just learned about in freshman chemistry./nOK. Good luck on the exam on Wednesday.
Tags // Transition Metals
Added: April 16, 2009, 11:36 pm
Runtime: 2982.13 | Views: 12897 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - Kinetics - Lecture 31
Principles of Chemical Science/n * Email this page/nVideo Lectures - Lecture 31/nTopics covered: /nKinetics/nInstructor: /nProf. Catherine Drennan
Tags // Kinetics
Added: April 16, 2009, 11:37 pm
Runtime: 2421.20 | Views: 10349 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - Transition Metals: Crystal Field Theory
Principles of Chemical Science/n * Email this page/nVideo Lectures - Lecture 28/nTopics covered: /nTransition Metals: Crystal Field Theory/nInstructor: /nProf. Catherine Drennan
Tags // Transition Metals Crystal Field Theory
Added: April 16, 2009, 11:37 pm
Runtime: 2422.93 | Views: 11837 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science -Transition Metals - Lecture 27
Principles of Chemical Science/n * Email this page/nVideo Lectures - Lecture 27/nTopics covered: /nTransition Metals/nInstructor: /nProf. Catherine Drennan/nTranscript - Lecture 27/nWe finished the unit on oxidation-reduction on Wednesday./nAnd so now we are moving into the next unit, which is transition metals, and that is Chapter 16 in your book. And following transition metals, the last unit is on kinetics or rates of reaction, and that is a little bit of a longer unit./nSo there are only two more units coming up in the course. Today I am going to introduce you to the transition metal unit, and this follows nicely from oxidation-reduction since you saw a lot of electrons being lost and gained./nAnd metals are definitely something that will gain and lose electrons, so it follows nicely from that./nToday is an introduction, and on Monday we are going to talk about crystal field theory, which is a topic that most people have not seen in high school. At least that has been true in past years./nIt looks a little bit scary sometimes to people because it looks different./nIt is not something that seems familiar. It is actually not that hard, but I strongly recommend coming and being present in lecture and working the problems. That is how you learn how to do this part./nAnd this often plays, this particular unit, on transition metals is a dominant theme on the final exam, so you will see it there./nIf people could settle down a little, I am even having trouble hearing myself./nThat is going to be Monday. And then on Wednesday we are going to talk about shapes of coordination complexes, and this fits in nicely with the transition metals. We are going to talk about vesper theory./nAnd then the following Monday we continue on and talk about how one can use crystal field theory to think about what colors molecules with transition metals have./nThese units, it is a little bit of a strange order this year because of the Thanksgiving holidays, but that is what is coming up./nAgain, this is a unit that most people have not seen before in high school. And, as I mentioned before, problem set nine, which is handed out today, you will see that it is really quite short. It just covers today's lecture and Monday's lecture./nIt just has a few problems on it, so it should just kind of keep you from falling behind with the material./nBut not take too much time so you will be able to get that done before Thanksgiving. All right./nToday there are a lot of topics. That is what happens when you introduce a new area. You need to go through a bunch of different kinds of definitions. And so today we are going to talk about coordination complexes./nAnd we are going to talk about coordination complex notation, coordination number, structures, just briefly of coordination complexes./nAgain, on Wednesday, we are going to be coming back next Wednesday and talking much more about these structures, but we will give a little introduction today./nWe are going to introduce something called the chelate effect and talk about isomers. And also, since we are talking about transition metals, we are going to start talking about d-electrons and d-orbitals./nThis is, again, all in Chapter 16 in your book./nHere is that relevant part of the periodic table that we will be talking about in this unit. These are what are called transition metals. And this is a part of the periodic table that is near and dear to the hearts of many faculty members in the Chemistry Department, including my own./nWe study proteins that use these metals to catalyze reactions./nAnd so this color scheme is actually out of a book that was published by Steve Lippard who is a faculty member in the Chemistry Department./nI will mention him again later in the day. He does a lot of work on the area called Metals in Biology, these transition metals that play a role in biological systems. And this book was also with Jeremy Berg who is the current head of the general medicine component of the National Institutes of Health./nAnd so he is somebody else who has thought a lot about metals in biology./nIn his case, in particular, zinc. Why are metals important in biological systems? Well, proteins, and I have talked some about proteins already, are composed of amino acids. You have carbon, nitrogen, oxygen, sulfur./nAnd there is quite a bit of chemistry that these particular elements can do, that amino acids can do, but there are some pretty challenging chemical reactions./nWe talked in the unit on chemical equilibrium about a challenging reaction, splitting nitrogen./nAnd I told you about the Haber-Bosch and how we need to do this to get fertilizers. And so it is an important industrial process that an enzyme can also do very efficiently without the high pressures and temperatures that human beings seem to need to get this reaction to go./nThat is one example of a protein that uses metals to do the chemistry./nAnd the active site of that enzyme is full of different kinds of metals. Up here says global cycling of nitrogen, hydrogen and carbon./nAnd so often some of the most fundamental reactions are also very challenging reactions and require metals to make them go. Radical chemistry reactions, making various products like vitamins or deoxy nucleotides, the building block of DNA./nAnd we have some DNA here with us today in class./nA lot of these reactions require metals. In this orange color here, those are metals that are found in biological systems, but there are also some metals in gray. And some have both orange and gray./nThere are other metals that are used as drugs or to probe biological systems./nThere is a whole area of sort of metals in medicine where people are using, say, looking at the use of gold for treating arthritis./nThere is a lot of relevance of these transition metals or d-block metals to medicine. Let's talk about the abilities of these metals to form complexes, and this is important for their ability to function in all systems, including the human body./nA key feature is that they can form complexes, and let's look at why./nAnd so these complexes then, of the metal and their ligands, are known as these coordination complexes. You have a positive metal ion, and it can attract electron density. And that is usually in the form of a lone pair of electrons from another atom./nAnd so when you bring together this metal and these other atoms and their lone pairs then you form this kind of complex./nAnd so if we look at the different roles that things play, you have a donor atom./nAnd they are called ligands. You often talk about a ligand being coordinated to a metal. And we mentioned Lewis acids and Lewis bases when we were talking about acid base, and we didn't really do much with them when we were talking about acid base./nAnd I said this would come back later./nLewis acids and Lewis bases are very relevant in terms of talking about how coordination complexes are formed. A Lewis base is going to donate lone pair electrons. And here are some examples. Again, these are referred to as ligands./nAnd so they would be acting as Lewis bases typically by donating their lone pair. Here are some of the ones that you are going to see in this unit./nAnd some of these will become familiar to you as we go through the next few lectures./nThey all have these lone pairs. And the metal is very happy to see ligands coming, and they will coordinate and form these coordination complexes. Then you need an acceptor atom, and that is the transition metal./nThat is going to act as the Lewis acid, and it is going to accept that lone pair that is being donated by the ligand./nLet's look at some example of some of these metals. And we saw some of these on the Periodic Table./nThese are all these d-block or transition metals. Then the coordination complex is the metal surrounded by the ligands or the Lewis acids surrounded by the Lewis bases. Here is a picture./nWe have cobalt in the center here, and then there are six NH3 groups which are acting as the ligands, and their lone pairs are pointing toward the metal./nAnd so, again, the coordination complex, all is meant by that is the metal surrounded by the ligands. And, in this case, there is a little bracket on the side with a plus three charge. And, for this particular complex, you might have like three chloride ions hanging around./nThree negative charged ions to kind of counterbalance the fact that this coordination complex has a plus three overall charge./nAnd so the charge would be distributed, but it is indicated that this whole complex has this charge by that little bracket and the plus three. Again here cobalt is acting as the Lewis acid. It is accepting the electrons that are being donated by the ligands./nIt is the acceptor atom./nAnd the NA3 groups are acting as the Lewis bases. They are the donor atoms, so they are donating a lone pair of electrons. Now we can introduce a term called coordination number, CN, and that is pretty simple./nIt is just a number of ligands that are bonded to the metal./nAnd so here there are six. The coordination number is six. That would be written as CN=6. And you have six ligands comprising the primary coordination sphere./nWe also have three chlorines hanging around, but they are not coordinated to the metal so they don't get counted in this. You are only interested in the ligands that are actually coordinated in the complex./nCN numbers typically range from 2 to 12, but 6 is the most common./nYou are going to be seeing a lot of different coordination complexes like this. Let me just show you how that would be written. When you are doing a problem in your book, it is mostly likely not going to be drawn./nYou are going to see notation for this instead. And so this particular one would be indicated, you have the cobalt and then you have your NH3 groups. There are six of them, so that is indicated there./nThen you have a bracket around the outside with a plus three, so that indicates that the charge of everything inside equals plus three./nYou also might see this where outside the bracket you have three chloride ions. And so you should be able to think backwards to this and say, well, if we have three things with a negative one charge./nAnd you see why the oxidation-reduction unit was important before this./nBecause we talked about oxidation numbers and all the rules for assigning oxidation numbers. You are going to need that knowledge to do this unit as well./nYou should know that this is going to be minus one for each chloride, so that would three, which would mean that if this is minus three, this bracket would be plus three, for this one down here is written to be neutral./nYou will be seeing this and need to back calculate./nThen you are also going to need to calculate what the oxidation numbers of the things within the bracket are as well. And we will talk more about that a little later in the lecture. Again, the NH3 within the brackets means it is bound to the cobalt./nThat would be this picture here. There are six of them so you would draw them as six around. And let me just mention a little bit about this particular configuration./nHere, where there is sort of a thick line coming out, indicates that it is coming toward you./nHere in this case you would have two atoms coming toward you. The dashed lines here are showing that there are two atoms back behind the plane. And then when you have the straight line, that is in the plane./nThis picture here is showing you this type of geometry, which is called what?/nOctahedral geometry, right./nLet's talk about octahedral geometry and the other geometries that you are going to see most often in this particular unit./nAgain, I am going to do much more on this on Wednesday, but this is just a very brief introduction to the structures that you will see the most in this particular unit./nFor a coordination number of six, you are just going to see one type of geometry, which is this octahedral geometry, which I just showed you./nFor coordination number of five, you will see two different types of geometry in this unit. And let's take a look at this. The first one over here, the trigonal bipyramidal./nHere is an example of that./nAnd I will try to have this as it is. Three ligands are in the plane. You have up and down and out here. One ligand is pointing out toward you with the thick line and another ligand is going back with the dashed line here./nThat is one there. We talked about the geometry./nAnd I forgot to mention that those are 90 degree angles, which I think is pretty easy to see when you look at it, that you would have 90 degrees in between those bonds./nAnd what about for this one? What do we have going for the red? What kind of angle is that? 120. And what about from the red to the black? 90. Those are the angles we would see and hear. And this is going to become important because we are going to be talking about d-orbitals./nAnd where the d-orbitals are for the metal, which is in the center of the coordination complex./nOne of the issues we will be discussing is where are the atoms with respect to those d-orbitals? Geometry becomes important at this point. The other type of coordination complex you might see, here the metal is gray in the middle, and so then we have our ligands bonded to it./nIf you have five ligands you also might see this geometry here, so this square pyramidal geometry./nAnd what are the angles between all of the atoms here? 90. Now, in the case where you have four ligands bound to your central metal there are two possibilities for the geometry that you will see./nOne is square planner. In this case, it is drawn with two ligands coming out towards you and two ligands going behind. And what would be the geometry between these guys?/n90 as well. And so this is one of the easy ones to remember because it is a square and it is planar./nThat is an easy one to remember. The other that you might see when you have four ligands is this guy tetrahedral. And what are the angles here? 109.5./nAnd then if you have three ligands, pretty much this is all you are going to see in this unit, and so we have the trigonal planar./nAgain, it looks trigonal and it is also quite planar. And what are our angles here? 120. And linear is going to give you what angle? 180. You will see this one as well./nAnd this model is a little old so it is not really so linear anymore but it is still supposed to be./nWe are going to bring back, I am going to bring these models to class for a lot of the next lectures because thinking about the structure is going to be very important in thinking about the properties of transition metals./nAll right. What else can transition metals do? What are some of the other things?/nWhen you talk about transition metals and coordination complexes, you need to talk about the chelate effect. We have seen so far ligands binding at one site, but some ligands can bind a metal at more than one site./nAnd so then you start talking about chelates. First a ligand that can bind a metal at one site is called unidentate or monodentate. And the dent comes from dentist or tooth./nThis would be one. If you have a ligand that can bind at more than one point of attachment then it is called a chelating ligand./nAnd the coordination complex that it forms is called a chelate. And so this is Greek for claws. It seems like claws are coming down and binding to the metal. Anything above one would be considered a chelate if it has two points of attachment or three or four./nBidendate means two points./nAnd this nomenclature, not too challenging. People are usually pretty good at guessing what they mean. Tridentate would have how many points of attachment? Three. Tetradentate? Four. Hexadendate? Six./nThere are some people, this was a question, I think, hexadentate on a final one year people got it wrong. Don't get something like that wrong./nSave the things you are going to get wrong for the really hard part./nThis is really not so difficult. That should be some bonus points there for you. If you remember the dentate, and this term actually gets used quite a bit, you will be all set. What is the chelate effect? Chelates are ligands that can bind at more than one point of attachment./nWhat is the chelate effect? Well, it turns out that metal chelates are very, very stable./nAnd so this is a thermodynamic property that they have. And it is believed, or the rationale for the stability which is observed, is that it is an entropic factor that is going on./nThe binding of that chelate at more than one point of attachment is entropically favorable because you are going to be releasing waters as you do this. We are going to come back, and I am going to explain this using an example in a few minutes./nFirst let me give you some examples of chelates and then I will show you how this chelate effect could work./nThere are examples of chelates in biological systems. And it may not be a surprise to some of you that one of the chelates is vitamin B12. And that would be one that I would pick to mention in particular./nHere the chelate, this is all one ligand. This whole bottom part is attached. It is one ligand./nWe have this called the corn ring. You have cobalt in the middle. And it is coordinated by this planar part./nAnd you have four nitrogen ligands coordinating. It is one ring system here called the corn ring that is attaching at four points of attachment. You also have an upper ligand here, and this is an adenosyl moiety./nAnd we have a lower ligand. And we talked about this before with buffers effects./nBecause, if this lower ligand is detached on the protein, it has to fit into a binding site. And genetic defects that block that impair the activity of the enzyme./nThat was mentioned before. This actually has six different ligands and will classify as a chelate or six different points of attachment and classifies as a chelate. Let me just sort of spin that around so you can kind of get a sense of the geometry here./nAgain, here is the cobalt in the middle, the upper ligand and the lower ligand./nAnd if we move that around, see if you can get a sense of the sort of geometry around that cobalt. You have four in a plane here, one on top and one on bottom, so it is sort of an octahedral type coordination of that metal in the center./nI thought I would just mention the person who determined the structure of vitamin B12./nThat is Dorothy Hodgkin. For her work on this, and on penicillin, she received a Nobel Prize, because at the time she was studying the structure it was really considered very large for the technique that she used, which is x-ray crystallography./nAnd people didn't think she would be able to determine where all those atoms were located in something that was that big using crystallography. Since then we have gone on to solve much larger structures./nBut she really opened the doors for people thinking about this technique as being important in biological systems./nShe had a number of graduate students who went on to do some pretty spectacular things and really helped to establish the field of crystallography. And I will just mention one of her graduate students who was not as successful as a crystallographer./nThis is what I do. And not everybody can do x-ray crystallography. It is a bit challenging./nAnd some people had to sort of find something else to do with their lives. One of her students didn't quite cut it as a crystallographer, but the good news is she was still able to find a job./nAnd it is actually kind of interesting that Dorothy Hodgkin was a real socialist. She was very involved in left-wing issues, but she stayed friend with Margaret Thatcher throughout the years. And it would have been very interesting to be in a living room with the two of them hearing them talk about politics./nMaybe they would talk about crystallography, I don't really know, if they were together./nThat might have been safer. That is one example of a chelate. Here is another example of a chelate, and this is called EDTA. And so it has many points for attachment to a metal. We have a lone pair up here on this atom, so we have a donor here./nWe have another lone pair, another donor site here. Another lone pair, another donor site here./nAlso over here and here and here. This ligand can attach to something in six places. Here, again, is the free molecule./nAnd here is what it looks like when it is attached to the metal. See this oxygen here, over here, and then if you follow it down over here this nitrogen is also coordinated. If we go over here then this oxygen is coordinated./nIf we come from this nitrogen over to the other nitrogen it is over here./nGo down to this oxygen, this oxygen is coordinated down here, and then in purple we have the last coordination. What is the geometry around that metal site? You have these six ligands, so what is that going to be?/nIt is going to be an octahedral./nYou should have angles around 90. And it is also hexadentate, so you have a ligand binding with six points of attachment. It would also have a CN number then of six. Why the chelate effect? Why are complexes like that so stable? The chelate effect could be viewed in this slide./nBefore you have the EDTA bound we have six water molecules around it./nThen we bind one molecule of EDTA. It binds in the hexadentate complex releasing all six waters. Where is entropy greater? When you have one free thing or when you have six free things? Six./nThis is entropically favorable. You are creating disorder./nYou are binding one thing and releasing six things. That is why these complexes are really quite stable. We are getting back to entropy now./nWe have been talking about delta G. We had delta H. Now we have delta S. That is the chelate effect. And we talk about it due to this entropic effect accompanied by the release of non-chelating ligands, usually water, from the coordination sphere./nIn this case, one ligand binding displacing six things is favorable./nThe chelate effect leads to a lot of practical uses. Here some of the uses would be in medicine. If you discover over Thanksgiving that you one and a half year old niece or nephew has discovered paint on the windowsill as peeling and is starting eating the paint off of the windowsill, you might want to remember when that was painted last and if there was lead in that paint and rush the child to the emergency room./nAnd when you get there, just in case the people there, you know, on Thanksgiving it is always dangerous to be in an emergency room./nYou might say give this kid EDTA. That is really what is used. And that will chelate then the lead out of the system before it can bind to things and do a lot of damage./nThis is one of the chelators that is used./nIt is a chelate effect. It works pretty well if someone has been exposed to something like lead, which is pretty toxic and you want to get it out of their system. That is one use. It is also in food./nIt is a food additive. Sometimes it is fun after you take a couple chemistry courses to actually start reading the labels of the food that you are consuming and some of those words will actually start making sense./nOne thing that you will often see is EDTA./nAnd it will say EDTA was added for freshness. One thing that EDTA can do. Microorganisms, bacteria, other things that you would rather not have consuming your food at the time that you also want to consume it need metals for life, as we all do./nIf you add EDTA, it is a good way of preventing other kinds of unwanted microorganisms from growing on the food./nBecause it will chelate all the metals and the organisms cannot live. Added for freshness is a more polite way of saying that this food is less likely to be contaminated with lots of bacteria growing on top of it./nAnother thing that it can be used for is bathtub cleaning or other kinds, but bathtubs and showers in particular. There is calcium in tub scum./nAnd most kinds of cleaners have something that chelates out that calcium which allows you to clean./nAgain, over Thanksgiving, one thing that happens to a lot of people when they go home is that their parents say help us clean to get ready for everyone coming. And so you might be using EDTA then as well./nI thought I would tell you a little story about how knowledge from freshman chemistry can make somebody money./nThis is a really kind of fun story. I don't know if it was around Thanksgiving, it might have been./nOne day Robert Black had a life altering experience. His wife said to him could you clean the tub? Apparently Robert Black had never cleaned the tub before that time, and so he went to clean it. Apparently, maybe it had not been cleaned in a while./nAnd he discovered that it is really hard./nYou have to scrub a lot to get the tub clean. He thought wouldn't it be nice if one never had to clean a tub? He came up with this idea. And here is one of the products that came out of this, Sparkling Shower, available at Leveret's./nAnd what he did was he put the same ingredients, including EDTA, in that you have in most cleaners, but he had sort of a novel pitch./nEvery time you take a shower, just squirt the shower or the tub, and then you never really have to clean it./nIn this patent he had a surfactant chelating agent, EDTA and an alcohol. You had a surfactant to sort of break up the water tension, the chelating agent to sequester out the salts in the scum and alcohol to dissolve the more oily ingredients./nAnd so those together kind of got everything loose and ready to go./nAnd the surfactant kind of helped it run off. If you squirt your shower every time you shower, you will never really have to clean it./nHe used all ingredients that were know before but packaged it slightly differently. Again, EDTA is one of the ingredients in these products. And so sales of these particular products have reached $70 million every year./nI think that the wife is probably very happy that one day she insisted that he clean the tub./nBecause this turned out to be a pretty profitable exercise. It is also an example of how some very simple chemistry can yield you quite a bit of money./nAll right. Tub cleaner. There was one more that I thought of. Anybody a fan of vampire movies?/nNot really all that many. It is a little surprising. I am not sure you are all telling the truth./nIf any of you happened to see Blade, you might be aware of another use for EDTA. If you remember, you may not have noticed what they were mentioning at the time, but they came up with a great way of killing vampires./nThere were these darts filled with EDTA./nI think the kind of general premise that you can rationalize this Hollywood special feature is that vampires drink blood, blood has iron, and if they are pretty much all blood, if you give them something that chelates out the iron it could cause some kind of harmful effect./nIn the movie they exploded and sort of disintegrated completely. If you take your niece or nephew to the emergency room for eating lead paint this will not happen./nI am quite sure. But it was really very spectacular./nHere they are, the little darts filled with EDTA. Again, more knowledge that could be useful on some occasion, perhaps, I am not really sure. I have a feeling that most of you will find the top three more useful, but here is the complete list that I have been able to come up with for the uses of chelators./nAll right./nThose are some coordination complexes, either single ligands bound around a central metal or chelating ligands bound around a central metal. There are other coordination complexes which have important features./nAnd one thing that we will talk about is isomers. Geometric isomers can have very different properties to them. It seems like it is pretty much almost the exact same compound./nBut the arrangement of the atoms is slightly different and that can yield two very different properties for those particular compounds./nLet's look at one example. This is platinum, and it has two NH3 ligands and two chloride ligands. And so that is in the bracket. Those four different ligands are coordinated centrally to the platinum./nIt would have a CN number of four./nAnd the most common shape that you would see for this would be this one here, this square planar shape. Depending on how you arrange the four different ligands two NH3 and two Cl, you get a very different result./nWe have, in this arrangement, a compound called cisplatinum./nYou have a cis arrangement where the chlorides are on one side and the ammonia groups are on the other side, or you can have transplatinum where the ligands are across from each other. Here are these two different molecules./nHere they are cis with respect to each other and here they are trans with respect to each other./nOne is a very potent drug that is used in treatment of cancer. The other has no known function. They are composed of identical atoms./nWhy is this different? Why are there different functions with the different geometries? Well, as it turns out, it is real important to have the atoms on the same side./nAs shown in this slide, I said I would mention Steve Lippard again, and here I am, this has been a subject of research in Steve Lippard's laboratory here in the Chemistry Department at MIT for a number of years./nIf you have a cancer cell and you want to kill the cancer cell, one thing that you can do is give it cisplantinum. The cisplatinum comes and binds to DNA. Here is a little cartoon of DNA here./nThere is the actual model of DNA structure down here in the front of the room./nWhen it binds then it can inhibit transcription. And so this can then lead to cell death. And so Steve has been really looking at this issue trying to redesign cisplatinum to see if you get a better drug, looking at other factors which might have an additive effect./nThis particular drug is used often for prostate cancer and has really been quite effective./nWhy does it matter if they are on the same side? Well, if you look at this complex and the other, you see that only two of these ligands are left./nTwo of the ligands are gone. At those sites it is actually coordinated to the DNA itself. That is why it is important that they are on the same side. You want to release two chlorides and have it bind to DNA this way./nIf it is on the other side, it really doesn't work as well. If we go over and look at the DNA molecule for a minute, you see that if you want to have two free sites to coordinate two DNA molecules./nOf course, this is the totally wrong scale./nYou want these on the same side. If I made it on the right scale it would not fit in the room, really. This is more or less, you can imagine. If you are going down here and you want to release just the chloride ligands, well, you cannot get the chloride ligands to be on the right side for this./nMolecules that have different shapes may have different properties./nAnd if you are in an environment you are looking to react with a particular molecule. I will show you another kind of view of DNA./nDNA is a really beautiful molecule. Here you have something you can carry around and show your friends, but also there is a smaller model of DNA. This is actually an artist in Michigan who will carve out within these blocks any atomic structure that you have./nAnd you can show the sort of bases in the detail./nThat is what DNA looks like sort of from the top and then from the side. It is just a really beautiful molecule. This is a nice example. Steve Lippard often takes a lot of UROP students, so it might be something, if you are interested in this area, that you might want to look at his Web page and see if that type of research interests you./nI lost my pointer./nThe cisplantinum has the effect because you need to release chlorides on the same side so that it can bind DNA over here. This cannot bind DNA because you don't have the chlorides together so they cannot be released and you won't interact with the DNA./nAgain, these different isomers have really, really different properties. Here we look at optical isomers or enantiomers./nAnd introduce the idea of a mirror plane here. When you have two different isomers, and if they are not superimposable, that means there is a mirror plane and that they are chiral./nAnd chiral molecules can have vastly different properties, again, if they are in a chiral environment such as the human body. Let's just look at these two./nSee that there is a mirror plane. Here we have the green on top, the NH3 on top, chloride in gray here on the bottom, and then on the side we have the oxygens, the black atoms here, more chlorides here, on the other side there are also oxygens pointing out in front, another NH3 group sort of putting in back./nAnd so they are very, very similar./nBut if you try to superimpose them, you will realize that it doesn't match up. If you try again other ways, put another green over here, no, that is wrong, too. And you can spend maybe not very long convincing yourself that, in fact, these are non-superimposable./nThey are mirror images so they are chiral compounds. And the idea of chiral compounds, we will come back, and you will hear much more about that if you take organic chemistry 5.12 or 5.13./nNow we have to get to some business that we need to think more about the properties of these elements./nAnd that is we need to count. A lot of counting in these units, oxidation-reduction, and also in this unit counting skills. We need to figure out how many d-electrons are involved in coordination complexes to be able to rationalize their properties./nThis is pretty straightforward./nOne can figure this out, what the d-electron count is if you know the group number from the periodic table and the oxidation number of the metal. We know how to calculate oxidation numbers at this point and the Periodic Table in your book./nLet's just take a look at a couple of examples and count some d-electrons. Suppose we have a coordination complex./nIt has cobalt in it, it has six ammonia ligands, and it has a charge of plus three for the overall complex./nThen we need to figure out what the oxidation number of cobalt is first. What is the overall charge on NH3? That is zero. This is a neutral molecule. And if you don't know that, this is something you will memorize as you go through this unit./nWhat does that leave you for cobalt? Plus three, because it all has to add up, just as it has before./nAnd so we figure out the oxidation number first. And then second we would figure out the d-count. We need to look at the Periodic Table and see where it is. Here we have cobalt in Group 9./nWe have Group #9 minus the oxidation number, which is three that we just calculated, and we are left with six./nAnd you will often see this written as d6 for the d-count number of electrons. One year when I had this, I think the first year, I actually had, I think, nine minus three equals five in the handout./nAnd that was rather embarrassing./nOh, wait. I want to do one more example first, or two more examples first. Let's try just a couple of more. All right. There are some that will become very familiar to you as you do these. Let's look at nickel with CO ligands, four of them, written like this./nThere is nothing up here./nThat means that the overall charge has to equal zero. CO, that is one that you will come across, and its overall charge is also going to be zero. It is neutral as well. And so what does that leave for nickel? Zero./nSometimes not very exciting. All right./nIf your oxidation number is zero then we want to look up and figure out what the d-count is. We have to see where nickel is. It is in Group 10, so we have ten minus zero./nIt is a d10 system. Let's just try one more./nCobalt, again, two molecules of water and one of NH3 and three of chloride to the minus one. Let's put that up a little bit. All right. Overall, this has to be minus one./nWhat is the charge on chloride going to be? It will be minus one./nThere will be three of them. NH3 was what again? Zero. Do you want to guess what water is going to be? Zero. Two zeros. What does that leave you for cobalt? Plus two. All right. Cobalt we looked up before and it was in 9./nOver here we have plus two./nAnd we have Group 9 minus two is seven, and so that is a d7 system. That is how you will do those problems. And you will need that on Monday when we start talking about crystal field theory./nIn the last couple of minutes, I want to remind you of the shapes of d-orbitals. This is leading into what we knew next./nAgain, we are thinking about the shapes of the d-orbitals and the metals, and then we are thinking about the shapes of the coordination complexes./nAnd we are trying to figure out whether the ligands are going to be near the d-orbitals or away. And depending on how the ligands are arranged with respect to the d-orbitals changes the properties of the coordination complex./nAll right. There are five d-orbitals. And you should be able to draw them./nAgain, the bar for successfully drawing the d-orbitals is you have to do as well as I do when I try to draw the d-orbitals./nAgain, it is a low bar for drawing. You just kind of need to get some of the general features right. If we look at dz2 it has a maximum amplitude along the z-axis. And in this course we will always try to keep the axes more or less the same to help you out./nAnd they will be labeled so you know where the z-axis is. Here is the z-axis and y is along the plane here./nAnd x is coming out toward you. There is x. And then you have a doughnut in the xy plane./ndx2 minus y2 has maximum amplitudes along x and y. It doesn't have anything along z. The orbitals are along y in the plane of the board and along x coming out toward you and also going behind the screen./nThen we have three more./nWe have dyz and it has its amplitudes 45 degrees off the y and the z-axis. Here is z up and down and y across the plane of the board. And so these orbitals are drawn to be 45 degrees off those axes./ndxy is 45 degrees to the X and the Z, so these are coming out towards you and going behind and also going up./nAnd then dxy is 45 degrees off of y and x. And you will need to know how to draw these and need to know the shapes and to be able to think about them in 3d so that you can imagine where ligands are and how those ligands are affected by these d-orbitals./nHave a great weekend.
Tags // Transition Metals
Added: April 16, 2009, 11:38 pm
Runtime: 3039.40 | Views: 17142 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - Kinetics (cont.) - Lecture 32
Principles of Chemical Science/nVideo Lectures - Lecture 32/nTopics covered: /nKinetics (cont.)/nInstructor: /nProf. Catherine Drennan
Tags // Kinetics
Added: April 16, 2009, 11:42 pm
Runtime: 2857.07 | Views: 6521 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Chemical Science - Kinetics (cont.) - Lecture 33
Principles of Chemical Science/nVideo Lectures - Lecture 33/nTopics covered: /nKinetics (cont.)/nInstructor: /nProf. Catherine Drennan/nTranscript - Lecture 33/nWe are getting close./nWe are in the last unit. We are doing kinetics. And we are going to talk about reaction mechanisms today. We had a little bit of an introduction to that already last lecture setting up to do this./nToday we are going to run through and talk about how you come up with reaction mechanism, how you evaluate a proposed mechanism in terms of the experimental data. This is, again, in Chapter 13. All right./nSuppose you have a particular reaction. You know that 2NO gas plus O2 gas go to 2NO2 gas. And someone has come up with some experimental data for you. And they figured out that the rate law for this reaction, there is observed rate constant, so K observed./nAnd in terms of the NO concentration, it is second order. And in terms of O2, it is first order. And so now you can come up with a reaction mechanism that fits that experimental data. And, as I mentioned last time, you never know for sure that your reaction mechanism is correct./nAll you can do is be consistent with the data. You can prove reaction mechanism is wrong. You can come up with a mechanism and show that that cannot possibly be the mechanism given the experimental data./nYou can prove things wrong, but you can never be sure that you have it right. The best you can do is consistent, but you can prove a lot of things wrong. And if you have a lot of mechanisms and you prove a lot of them wrong and there is only one that seems to hold then there is a good chance that probably is correct, but you never 100% because someone might do another experiment and discover something else./nBut, in the problem set, you will be saying, yes, this mechanism is consistent, or you will be writing a consistent mechanism, or you will be saying that mechanism cannot possibly be true. All right./nLet's look at that. First let's just consider what the overall order of this reaction is. What is the overall order if this is the rate law? Three. I am just running through these. We will practice some of these because these are good couple point questions on the final, and so you don't want to miss these easy points./nAll right. If the overall mechanism is three, the overall order is three, three things coming together at once to react. Is it likely that is it one step? No, it is not. What is it called if you have three things coming together in terms of molecularity? Termolecular./nAnd those are rare, as we talked about last time. As I said, if you are getting together three people, it is usually not true that they all arrive sort of ready to go at the same exact moment. Usually one person is late, if not more./nGetting three things to react together at the same time is challenging, and so those are rare. So, this probably has more than one step. Then we have to try to divide up this reaction into steps and come up with a mechanism, a two-step mechanism we will try for this./nLet me leave that up. All right. Let's draw a potential mechanism on the board. And then we will try to figure out, for that mechanism, what the rate law would be if that was the correct mechanism and see if it is consistent with the experimental data./nAgain, what was the experimental rate law is K observed. That is the observed rate constant. Second order in NO and first order in O2. We can write the reaction in two steps, and each step is an elementary reaction so it occurs exactly as written./nIn the first step, we will have two NOs coming together. And the forward rate constant will be k1 and the reverse rate constant will be k-1. And it will be forming an intermediate N2O2. In the second step of the reaction we will have oxygen come in and interact with the intermediate, N2O2, and form, with a rate constant k2, two NO2s./nAnd that will give us our overall reaction. Now we need to write rate laws for each step. And since each step is an elementary reaction, we can write the rate law using the stoichiometry of that step as written./nThe rate law for the forward reaction would be what? What do I put first? k1. And then? NO squared. What about the rate for the reverse reaction? k-1. Times what? N2O2. OK. Let's look at each of these./nFor the forward direction, what is the order? The overall order is two. And so in terms of molecularity what is it? Bimolecular. For the reverse reaction the order is what? One. And what is that called? Unimolecular./nOK. Let's look at Step 2. The rate for step two is what? k2. Times? It would be O2 and N2O2, the concentration of each. Again, we can write them exactly as written. Now we consider the rate of N2O2 formation, so the overall rate of NO2 formation from these two steps./nAnd we can just use second step where NO2 is being formed. And so we can write this is equal to that rate law times two. Let me rewrite the rate law. And the two comes from the fact that two NO2s are being formed./nBecause two NO2s are being formed, this concentration is increasing at a different rate than the concentrations of these are decreasing. The book is inconsistent with whether it puts a two all the time, so I am not going to take off if people forget to put the two in this./nBecause the book should be more consistent with that. But, if you do see a two, it will mean that you are forming two of the products in the reaction. All right. Now we have written a rate. We are not equal to our experimental rate here because there is an intermediate./nThis is an intermediate. And an intermediate cannot appear in the overall rate law for something. You are going to be spending a lot of time getting rid of intermediates. Let's now leave this here and try to get rid of our intermediate./nI am having trouble with papers today. Now let's solve for the intermediate. And we want to solve for this intermediate in terms of things that are not intermediates. We want to solve for it in terms of reactants or products or rate constants./nWe need to consider then how this intermediate N2O2 is being formed and how this intermediate is being consumed or decomposed. We want to figure out the net formation of the intermediate. And the net formation of the intermediate is going to equal the rate at which it is formed./nIt is being formed over here in the forward direction of the first step, so we will write that down. We consider how it is being formed. It is being formed with the rate k1 times the concentration of NO squared, so that is just the rate law for the forward direction of the first step./nIt is also being decomposed in the first step in the reverse direction. It is being decomposed with the rate of k-1 times the concentration of the intermediate. That is just our rate law for this reverse step of the first reaction./nAnd it is also being consumed. It is also being consumed in the second step with a rate law of k2 times the concentration of the intermediate times the concentration of O2. That is just the rate law that we wrote for this second step./nIt is being formed, decomposed and consumed. The net formation equals the rate at which it is formed minus the rate at which it decomposes minus the rate at which it gets consumed. And now we can use something called the steady-state approximation./nIn the steady-state approximation, all it is saying is things are in a steady state. That means that the net formation of this intermediate is equal to zero or, if it is in a steady state that the rate at which it is being formed is going to be equal to the rate at which it is being consumed, either by the reverse Step 1 or Step 2./nThis rate is equal to this rate. That means it is steady, so it is in the steady state. The net formation is zero. The rate at which it is being formed equals the two rates at which it is being consumed or decomposed, the rate at which it is decaying are equal to each other./nThat is the steady-state approximation, and it is a very good approximation. And so you can use this for all of the mechanisms that you will be writing. That allows us to set it equal to zero, or we can rearrange these terms so that these two are on one side and the rate at which it is formed is on the other side./nLet's do that. And I am going to take these two and put them together and also pull out the term for the concentration of the intermediate because I want to solve for that term. I am going to try to pull it out and solve for it./nIf we rearrange this then, we get that the concentration of our intermediate, you pull it out from those terms, so that leaves k-1 plus the other term k2 times O2. Now I have just pulled this out from these two, and I have moved these to the same side of the equation./nAnd so that is going to be equal to k1 times NO squared. This is the rate at which you are decaying your intermediate and this is the rate at which you are forming that intermediate. Now I can continue to solve for this./nThe concentration of the intermediate is going to be equal to k1[NO]2 over k-1 plus k2 times the concentration of O2. Here is what the intermediate then is equal to. And now I can take this term and go plug it back in./nAll right. We can take the concentration and plug it back into here, so that whole term that we just came up with. And we will put it back into that equation. Then, if we substitute, we get the rate of NO2 formation is going to be equal to two rate constant k1, rate constant k2 times NO squared times the concentration of O2 over k-1 plus k2 times the concentration of O2./nBut that is not consistent. We have NO squared and two O2 terms. Whereas, in our observed experimental law, there is only one term for O2. That is not consistent so something is wrong. This is not consistent so there must be fast and slow steps./nIf there are fast and slow steps then we can simplify this equation and make it like the experimental equation. Now we can go over here and think about what would happen if Step 1 was fast. It is a fast and it is a reversible step./nIt is already written that it is a reversible step, so it is reversible. And Step 2 might be slow. Now we can consider, if we have a fast first step and slow second step, what that would do to this equation and whether we could then rearrange this equation such that it would be equal to our experimental rate law./nWhen we say something is slow, we are going to introduce a new term. And that the slowest step in a reaction mechanism is called the rate determining step. And, for a rate determining step, you can assume that that step is governing the overall rate of the reaction./nIt is slow enough that the rates of the other reactions don't really matter. That is what rate determining means. And let me just give you an example of a rate determining step. After today's lecture, you will be able to do all of the problems on Problem Set 10./nThis is your last problem set in 5.111, and so it would be really great when that problem set is finished and turned in. And so all through the rest of this lecture thinking about I can go out right after class and finish that problem set because you are very excited./nAnd by the time it gets close to the end, you see the end of the handout coming up. Your books are sort of packed, you are ready to go, and then the minute it is finished you are out the door. And you are going to some place where you can sit down and finish Problem Set 10./nYou are moving really fast. You might even be running. You may be even jumping over people in the infinite corridor to get to the library as fast as you can. You get to the library and you look for a table, but everyone else from 5.111 is already there and all the tables are filled./nYou go to the first floor and the second floor and the third. You go to the basement and in the stakes. Every table is filled up with people doing Problem Set 10. Then you have to leave the library./nYou think Building 2, lots of small classrooms, they cannot all be in use. You run over to Building 2, but all those classrooms are filled with 5.111 students doing Problem Set 10. And so eventually you get to Building 4 and you find a free classroom./nThe minute, the second you find that free classroom your books are out of your bad. They are out. Your calculator. Your correct kind of approved calculator is right there and you are ready to go. It takes you maybe two seconds to get out [10 at 2:50?] and it takes you maybe two seconds to pull out all your books to be ready to go, but it takes you a half an hour to find the table to do that problem set./nOverall, half an hour plus four seconds is a half an hour. That was the rate determining step. In finishing Problem Set 10 was finding that table. What you can do for these as well, if we say there is a slow step, you can assume that that is rate determining./nAnd you don't have to worry about how fast those other steps are. It just depends on the rate of the slow step. That is governing the overall rate at which the reaction or the process happens. That is rate determining step./nWe can use that now. And we can use that and go back and simplify the expression that we solved for the intermediate. Here we were solving for this expression, and we weren't considering anything about fast and slow steps at this point./nAnd so this is the equation that you would get if there were no fast and slow steps. And then we plugged it into this and found that it was not consistent. Now let's take a step back and go back here and think about what this expression would be if the first step was fast and the second step was slow./nHere, again, I have just put it up on the screen what we had on the board. We have this first step is fast and reversible and the second step is slow. Really, what we are talking about here is we are comparing rates for the decomposition of the intermediate, that is the reverse direction of step one, with the rate for the consumption of the intermediate./nWhich of these is faster? Well, decomposition is faster because this is a fast step. And the consumption in Step 2 is slow. What does that mean? Well, if we write it down then if the rate of the decomposition, this rate, the rate of the reverse reaction is faster then this k-1 is a bigger number than k2./nThe rate of consumption here is slow. We are going to consider the relative size of these numbers and how they fit in this term, and you can drop things out by assuming that one thing is a lot bigger than the other thing./nIf we do this then we are saying, again, this is the same thing that was on the last slide, that the rate of the decomposition is faster than the rate of consumption. i.e., Step 2 is slow, Step 1 is fast./nWhat you are saying is that k-1 is a bigger number. That rate constant is a bigger value than the k2 term. And, in the bottom of the equation here, in the bottom of this equation for the intermediate, we have a term k-1 and we have the term k2 times the concentration of O2./nAnd now, if you are saying this about the different rates of this reaction, if you are saying this is fast and that is slow, this is a big number, that is a smaller number. If you have on the bottom here a really big number plus a much smaller number in comparison, this smaller number is not adding much to the bigger number./nAnd we can just cancel that term out. We can just say it is so much smaller than the other, we are not going to worry about it, and we are going to get rid of it. Now our expression would look like this./nThe concentration of the intermediate is k1 times NO squared over k-1. We can rearrange this and think about does this look familiar to any of you in terms of what we have seen before. We can rearrange it to look like this./nAnd what is this term? What is this all equal to? K, the equilibrium constant. It is equal to k1. This is just an equilibrium expression. And, if you go back and look over here, you will see that if you are going to write the equilibrium expression for Step 1, products over reactants, that is what you would get./nAnd also, we saw last time, that you can express equilibrium constants in terms of rate constants. And the equilibrium constant for a reaction is going to be equal to the rate constant for the forward reaction over the rate constant for the reverse reaction./nAnd at equilibrium those rates are equal to each other. So this is just an equilibrium expression. When we assumed Step 2 was slow and Step 1 was fast, we can now solve for the intermediate in terms of an equilibrium expression./nLet's consider why this is true. OK. When you have a fast first step followed by the slow step, the first step is pretty much in equilibrium. The first reversible step is in equilibrium. And you can kind of think about it in terms of this diagram./nHere you have reactants going to intermediates, and this is fast and reversible. Then the second step, the consumption of the intermediate is slow going to products. If this is really pretty slow and not much of the intermediate is being siphoned off to product, more of it in fast equilibrium is going back and forth with the reactant, then that creates sort of an equilibrium expression./nIt is not perfect. Some of it is being siphoned off, but more or less you can consider this in equilibrium. Again, if you have a fast reversible step followed by a slow step, you can assume equilibrium conditions./nThat is going to make your life easier in terms of solving these expressions, because you can go right to writing an equilibrium expression then for your intermediate. Now we can go back and substitute./nWe have this term now for solving for our intermediate. Or, you could write rate constant k1 over k-1. Or, you could write big K, the equilibrium constant k1. These are equal to the concentration of the intermediate./nWe can put this term back in here now. This was, again, what we wrote in the very beginning for the formation of NO2. And we can substitute that back in. And we can have it in terms of the rate constants or using an equilibrium constant in there./nThose would be equivalent. And all of these Ks, in terms of the experiment, are going to be grouped together as a K observed. You usually are not going to measure all the individual rate constants./nYou are just going to have some overall rate constant that you do measure, an observed rate constant. And so if you write then either of these in that form where all your Ks just get grouped into K observed then you see that that is consistent with the experimental data./nThat mechanism works. A mechanism works that has two steps, the first step being fast and reversible and the second step being slow. We don't know for sure that that is the mechanism, but it is consistent with the experimental data./nLet's look at another example now. We are going to look at three different examples. Here there is a proposed mechanism that is very similar to the one that we just looked at. Now we are going to try to do this problem in a more efficient way than we did last time, which is kind of work through it the way you will when you are doing Problem Set 10 really fast./nAll right. We have a two-step mechanism, one that is fast and reversible and a step that is slow. The first thing you want to do is write the rate laws for each of the individual steps. For this one, what is the rate law for the forward reaction? k1 times O3./nThat is for the forward reaction. For the reverse reaction we have k-1 times O2 times the concentration of this intermediate O. Let's look at the second one. Here we have the rate k2 times the intermediate O times O3./nNow the rate is going to be determined by this slow step. And we can write the equation for the formation of O2 from this last step. And we could do that whether it was the slow step or not. We did that before as well./nIf we write the formation of O2, it is being formed in the last step from O and O3. We are using this rate law. And, again, we have a two in there because, like the last example, two things are being formed./nWe have to molecules of O2 being formed. The rate of formation of O2, two times k2 times this intermediate times O3. If there is no intermediate in the expression we would be done, but there pretty much always is going to be an intermediate in the expression./nWe need to solve for that intermediate. O is an intermediate, so you need to solve for it in terms of reactants, products and rate constants. Again, here we have this step that we saw before. The first step is fast and reversible, the second step is slow, and so you can assume, approximate that the first step is in equilibrium./nAnd so you can solve for the intermediate in terms of an equilibrium expression. What is the equilibrium expression for the first step? K would be equal to? I will just put it down here. You can write it in terms of k1 over k-1, products O2 times O over reactants O3./nThat is just our equilibrium expression. We have it in terms of the rate constants and in terms of this expression for products over reactants. And now we can take this and rearrange it to solve for our intermediate./nWe would have this, we would bring the O2 down to the other side, we would bring the O3 up here, which leaves us with k1 times the concentration of ozone over k-1 times the concentration of oxygen gas./nThat was a lot simpler than what we did over here. And so you can feel free, if you are told there is a slow step and a fast step, just jump right to that and solve it in terms of the equilibrium expression./nIt will save you some time. You should come up with the same answer if you do it the way that we did it in the first step. And sometimes you will have to because you won't know about fast and slow steps./nNow we can take this term and substitute it back into the rate law for the formation of O2 gas. And so we had two k2 times this intermediate times O3. We put this term in for our intermediate and we get two times k2 times k1 ozone squared over k-1 times O2./nAnd so this would be the rate law for this particular mechanism. If that is the rate law for that mechanism, you could think about what are the orders in this and could you experimentally go back and look and say, OK, is that consistent with the mechanism? Let's take a look at what we would predict from this./nFirst we could also write this in terms of K observed and group all the K constants together for K observed, and that is times ozone to the two and O2. If we look at this, if we were going to design experiments for someone to do, what would we predict to sort of prove that that is, in fact, a correct mechanism? What is the order, based on this rate law, in terms of the concentration of O3? Two./nIf someone went into the lab and doubled the concentration of O3, what should they observe happen to the rate? They should observe four times the rate. What is the order in terms of O2? What is it? Minus one, right./nIf you doubled the concentration of O2, what should happen to the rate, keeping everything else the same? It should be half. This is the kind of thing that can be experimentally tested, whether it is consistent./nAnd what is the overall order then of this reaction? It would be one. And so if you doubled both things, what should you observe happen to the rate? Yeah, it should double. You could go back and sort of test this./nOK. Let's look at one more example. This time we are not going to make a proposal ahead of time about fast and slow steps, but we are going to work out what the overall rate law for this is without fast and slow steps and then go back and consider which step, or any, are slow./nNow we are going to go through the long way and do this problem and then think about what we can learn from it in terms of we are given an experimental rate law. And once we see how this works out, we can think about whether a step is slow or not./nFirst we start the same way, and we are going to write the rate laws for each of the individual steps. Again, they are elementary reactions. Steps in a mechanism are elementary reactions. We can write the rate law exactly as the equation is written./nWhat is the rate law for the forward direction here of Step 1? First we have k1 times NO concentration times BR2 concentration. That is the forward. And in the reverse we have rate constant k-1 times NOBR2./nFor step two, we have rate constant k2 times NO. There you go. So, you know how to do these. We have that. Now we are going to be looking for the overall formation of NOBR, and so we can write this just using the last step./nWe can use that rate expression for that last step. The formation of this. And we are putting a two in because, once again, there are two of these being formed in that last step. Again, we could just use the rate expression for the last step./nAnd there is a two in there because two are being formed. That is the rate of formation of NOBR. Now we have an intermediate. This is being formed and is being consumed, so we need to solve for the intermediate./nThere cannot be any intermediates in our overall expression, so we need to get rid of that. So, it is intermediate and we need to solve for it. Now, we cannot go and just use the equilibrium expression for the top because we don't know anything about fast and slow steps./nWe don't know that that is going to be a good assumption. Let's just write it out the long way and then come back and consider whether that is consistent with the experiment or we have to do some modifications./nWe are going to solve for this in terms of reactants and products in Ks. Now we want to consider the net formation or the change in the concentration of this intermediate. The intermediate is being formed in the first step, so we can write down the rate expression for how it is formed, which is just this, k1 times NO times BR2./nThis is how it is being formed. It is decaying in this second step, in the backwards step, so we can put that in. We have a minus sign here. Because it is going away by the rate expression of k-1 times the concentration of the intermediate and it is also being consumed in the second step./nIt is being used up in that second step, so we will put that term in here. That is minus k2 times the concentration of the intermediate times NO. It is being formed in one step and it is being consumed in two different steps./nNow we can use the steady state approximation which says that the overall net change is zero. We can set this term equal to zero. It is a steady state approximation. That is the same as saying that the rate at which an intermediate is formed is equal to the rate at which the intermediate goes away./nWe can rearrange these terms, bring these on one side of the equation and have this one on the other. Rearranging then we moved these onto one side and the formation to the other side. Now we can pull out the term for our intermediate./nTake it out here, that leaves k-1. Take it out here, that leaves k2 and NO. Again, this is the rate at which it is being formed. And now we can solve for the intermediate. When we solve for the intermediate, we take this term and divide it by that term and we get this expression./nThat is solving for the intermediate. The next step is to take that intermediate and plug it back into the initial equation that we had, the formation of NOBR, and that was two times k2 times the intermediate concentration times NO./nNow we take this whole term and put it back into this term and we get this expression here. All right. This expression now is not consistent with this experimental rate law, so there must be fast and slow steps./nNow let's consider what is going to happen if the first step is slow and the second step is fast. Again, you are considering what is happening on the bottom of this expression. You are asking how does k-1 compare to this k2NO term./nAnd so if you say that the first step is slow then k-1 is small. Second step is fast. That means k2NO is a bigger term. If this second step is fast, the rate constant for the second step is a bigger number than the rate constant for the first step./nAnd so what you can do then is you can say, OK, that is small compared to this so we drop out that term. If we drop out that term, can we cancel anything else? What else can we cancel? We can cancel k2./nWhat else? We can cancel out one of the NOs. And that leaves us with this expression here. There is the expression again, two times k1NOBR2. And that can be written as K observed. And, look at that, it is consistent with the experimental rate./nNow let's just check the other way to make sure one should be right and one shouldn't be right. That we can get the answer if we go the other way around. This is consistent with the experiment. And the overall number of this is two./nAll right. Now let's look the other way around. If the first step is fast and the second step was slow that would mean that k-1 would be a lot bigger than the k2 term. k-1 rate constant for the first step./nk2 rate constant for the second step. Fast. Bigger than this. And so then we can take this term out. Can we cancel anything else? No, you cannot cancel anything else. This is the expression. We can put that down here./nWe can group the K terms to K observed. And so this is not consistent with the experiment that you had before. That way doesn't work. You can write out the whole thing and then think about whether they are fast and slow steps./nAgain, you will be asking the question which of these terms is bigger than the other term? What is the fast step? How does k-1 compare to k2? You can make the simplification. And only one should be consistent./nOne mechanism should be consistent. The other one should be inconsistent. OK. Now you can do Problem Set 10.
Tags // Kinetics
Added: April 16, 2009, 11:43 pm
Runtime: 2566.87 | Views: 8694 | Comments:0

Type: public Video
Status: Live!

Share this video with friends! Copy and paste the link above to an email or website.

Share Video

My Tags